Next Article in Journal
Economic Innovation Caused by Digital Transformation and Impact on Social Systems
Previous Article in Journal
Impact of Pandemic COVID-19 on Air Quality at a Combustion Plant and Adjacent Areas
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Thermochemical Conversion of Biomass for Syngas Production: Current Status and Future Trends

1
Department of Chemical Engineering, Dawood University of Engineering and Technology, Karachi 74800, Sindh, Pakistan
2
Department of Environmental Sciences, Sindh Madressatul Islam University, Karachi 74000, Sindh, Pakistan
3
Department of Chemical Engineering, Quaid e Awam University of Engineering Science and Technology, Nawabshah 67480, Sindh, Pakistan
4
Asian Institute of Fashion Design, Iqra University, Karachi 75500, Sindh, Pakistan
5
Department of Energy and Environment Engineering, Dawood University of Engineering and Technology, Karachi 74800, Sindh, Pakistan
6
Key Laboratory of Pollution Process and Environmental Criteria, Ministry of Education, College of Environmental Science and Engineering, Nankai University, Tianjin 300350, China
7
Department of Chemical Engineering, Mehran University of Engineering and Technology, Jamshoro 76062, Sindh, Pakistan
*
Authors to whom correspondence should be addressed.
Sustainability 2022, 14(5), 2596; https://doi.org/10.3390/su14052596
Submission received: 8 January 2022 / Revised: 11 February 2022 / Accepted: 11 February 2022 / Published: 23 February 2022

Abstract

:
The thermochemical conversion of different feedstocks is a technology capable of reducing the amount of biowaste materials produced. In addition, the gasification of feedstock using steam as a gasifying agent also produces hydrogen, which is a clean energy fuel. This article aimed to encapsulate the current status of biowaste gasification and to explain, in detail, the advantages and limitations of gasification technologies. In this review paper, different gasifying agents such as steam, air, and oxygen, as well as their effects on the quality of syngas production, are discussed. In addition, the effects of reactor configuration and different operating parameters, such as temperature, pressure, equivalence ratio, and incorporation of a catalyst, as well as their effects on the ratio of H2/CO, LHV, syngas yield, and tar production, were critically evaluated. Although gasification is a sustainable and ecologically sound biomass utilization technology, tar formation is the main problem in the biomass gasification process. Tar can condense in the reactor, and clog and contaminate equipment. It has been shown that an optimized gasifier and a high-activity catalyst can effectively reduce tar formation. However, key biowaste treatment technologies and concepts must first be improved and demonstrated at the market level to increase stakeholder confidence. Gasification can be the driving force of this integration, effectively replacing fossil fuels with produced gas. In addition, support policies are usually needed to make the integration of biomass gasification technology into the industry profitable with fully functional gasification plants. Therefore, to address such issues, this study focused on addressing these issues and an overview of gasification concepts.

1. Introduction

Global warming and environmental problems linked to the extensive use of fossil fuels have diverted research toward renewable energy resources. Furthermore, nonrenewable energy resources are being depleted while the world’s energy demand continues to grow tremendously [1,2,3,4,5]. Renewable options include solar, wind, tidal, wave, hydro, and biomass energy. Among all existing renewable energy resources, the conversion of biomass for energy production contributes to over 70% of all renewable energy generation. The participation of biomass in the world’s primary energy supply accounts for over 10%. The application of biomass for bioenergy production offers many advantages. Bioenergy can be used in the transportation sector and for the generation of heat and electricity [4]. It increases the diversity of the energy supply [6]. Moreover, bioenergy can yield solid, liquid, and gaseous fuels. The energy obtained from biomass can be stored and transported to different sectors [6,7].
As biomass is enormously available throughout the world, the energy production from biomass is comparatively cheap compared to other energy feedstocks. Biomass is ecofriendly, as it absorbs part of the CO2 that is discharged during the consumption of fuel, and at the same time reduces greenhouse gas emissions [5]. The application of biomass for bioenergy will lead to waste management control and initiate socioeconomic and regional development where biomass resources are utilized for energy generation by providing income and jobs in rural areas. Even though the potential of bioenergy is obvious, more than half of bioenergy is collected through the application of traditionally used biomass, which consists of agricultural residues, animal dung, charcoal, and wood for heating and cooking purposes. The traditional route of biomass utilization results in low energy efficiency and hazardous emissions that cause health issues [6]. Therefore, in the development of sustainable bioenergy, the enhancement of energy efficiency and modern energy recovery practices have been adopted.
The modern biomass energy-conversion method includes thermochemical conversion and biochemical conversion routes for bioenergy production. Different microbes and enzymes are introduced to convert biomass into synthetic oils, biodiesel, biocrude, and bioalcohols [8,9]. The addition of enzymes and microbial cells in the biochemical conversion route makes the product costly, and is time consuming [10]. On the other hand, in thermochemical conversion pathways, heat and a catalyst are applied in the transformation of feedstock into intermediate products [8,11]; for example, bio-oil and syngas. Thermal treatment for the conversion of biomass is robust and flexible, since it can be used for different varieties of biomass as compared to biochemical conversion routes [8]. Among various thermochemical conversion routes of biomass, gasification is the eye-catching option, as it is not merely environmentally friendly, but its energy efficiency is higher than pyrolysis and the combustion products of biomass [4,12,13]. Therefore, the aim of this study was to present a state-of-the-art analysis of available biomass conversion technologies. Additionally, possible future trends are discussed regarding efficient energy recovery and applications of recovered biomass resources.
In gasification, biomass materials are converted into a combustible gas at a temperature ranging from 600 °C to 1500 °C using gasifying agents. The oxygen feeding is maintained below stoichiometric oxidation limits. When gasification is conducted by maintaining the temperature at lower levels, the gas produced, which is called product gas, may be composed of CO, CO2, H2, CH4, and low amounts of hydrocarbons [14]. Moreover, when gasification is performed at higher temperatures, the gas produced, which is called syngas, is mainly composed of H2, CO, CO2, H2O, and fewer contaminants [15]. The thermochemical and biochemical conversion pathways of biomass are shown in Figure 1.
The syngas obtained through gasification finds wide application in gas turbines and combustion engines, and in biocatalytic and catalytic processes for the synthesis of alcohols, hydrocarbons, organic acids, and esters. The cost competitiveness and performance data of syngas make it one of the most competitive renewable energy options. Although the syngas derived from biomass is more expensive than the syngas derived from coal, biomass-derived syngas is still a more competitive energy source than naphtha and diesel, making its research even more captivating.

1.1. Gasification History and Reactions

Gasification history can be traced back to the 16th century. The idea of gasification went through several stages of development. Among the various thermochemical processes for converting biomass into products, gasification is widely applied using a limited supply of oxygen or any other oxidizing agents including carbon dioxide, steam, or a mixture of both to transform solid biomass into liquid or gaseous biofuel. The gasification reactor is also called a gasifier [16,17]. Compared to combustion, the efficiency of converting biomass to biofuel through gasification is higher. Consequently, the biofuel achieved using the gasification process is a promising technology and provides better efficiency in energy generation. Biomass gasification plants range from a few kilowatts up to several megawatts of power generation [18,19]. Furthermore, the efficiency of the gasification plant ranges from 70% to 80% [20], depending on the gasification agent used. The main product obtained from biomass gasification is syngas, commonly known as synthesis gas. The resulting synthesis gas composition contains CO, H2, CO2, and CH4. These gases are burned in a gas turbine to generate electricity. Biochar is a by-product of biomass gasification that is utilized in the agriculture field to improve soil fertility [16,17]. The composition of the syngas changes with the gasification agent used during the gasification of biomass. The most frequently used gasifying agents are carbon dioxide, oxygen, steam, air, a mixture of air–steam, or a mixture of air and carbon dioxide streams in different proportions. Among various gasifying agents, the air is extensively applied due to its high availability and zero purchase cost [16]. Gasifiers are also designed to convert other biomass raw materials and the dry organic part of municipal solid waste into energy [21,22]. The types of gasifiers used for energy recovery from biomass [17] are mainly fixed bed, EFG, and FBG. Table S1 (Supplementary Information) shows a detailed review of biomass gasification. Table S1 presents a short review regarding the kinetic parameter estimations for different feedstocks. It has been observed that different mathematical models, including the volume reaction model, shrinking core model, random pore model, etc., were applied to estimate the kinetic parameters, i.e., pre-exponential factor or Arrhenius constant (A) and activation energy (E). Both isothermal and non-isothermal techniques were applied in the experiments. Due to the inconsistency of the dimension of A (change in order and mathematical expression used), the comparison of A seems to mean less. However, activation energy was found in the range of 30 to 243 KJ/mole for all the used feedstocks except the mixture of scrap tire powder, industrial waste, and sewage sludge for which the activation energy was estimated in the range of 65.7 × 103 to 185.4 × 103 KJ/mol. The possible reason for this high activation energy is the presence of higher hydrocarbons in scrap tires. It is a common understanding that a high amount of activation energy has a negative effect on combustion/gasification. Table 1 represents different fuels, heating values, and a comparison of prices. The major reactions during biomass gasification include endothermic and exothermic, such as partial oxidation, steam methane reforming, water gas shift reaction, Boudouard reaction, etc., as explained in Table 2.

1.2. Characteristics of Lignocellulosic Biomass

The generic term for non-fossilized organic matter and biodegradable material is biomass. It is produced directly through photosynthesis and is usually applied for the manufacturing of chemicals and fuels. Moreover, the most widely available biomass resources for energy production are wood, starch, and sugarcane. Generally, wheat, corn, cereals, oil crops (canola, sunflower, palm, rapeseed), and nonfood crops include eucalyptus, miscanthus, willow, etc. Non-lignocellulosic biomass residues from agricultural and forest residues include wet organic materials such as animal wastes, sewage sludge, and organic waste fractions from municipal solid wastes, and are commonly used for bioenergy production [23]. In bioenergy production, the more commonly used feedstocks are lignocellulosic material due to their low cost and abundance [25]. Table 3 shows the main components of biomass.
The composition of lignocellulosic biomass comprises proteins, lipids, extractives, and carbohydrates [7,27]. The carbohydrate in lignocellulosic biomass is up to 75%, which is particularly prominent in the form of hemicellulose (23–32%), cellulose (38–50%), and lignin (15–25%), [28]. Lignin, cellulose, and hemicellulose are intermeshed and physically connected by covalent and noncovalent bonds [28,29]. Figure 2 shows the intermeshed structure of lignocellulosic biomass having cellulose as its main component [27,30]. Cellulose is not soluble in most solvents and is not prone to enzymatic hydrolyzes and acids [8,25]. In comparison to cellulose, hemicellulose has some distinctions, as hemicellulose comprises polysaccharides other than glucose such as pentoses and hexoses [9,25]. The structure of hemicellulose is smaller in length, marginally branched, and not as crystalline as the structure of cellulose. Lignin has a three-dimensional configuration comprising aromatic monomer units, and it is a high molar mass molecule [27]. Lignin is composed of several types of arrangements that repeat unsystematically over cross-linked covalent bonds, and is a relatively hydrophobic polymer.

2. Effect of Feed Composition on Gasification

2.1. Composition of Feedstock

The potential of feedstock is determined through the investigation of proximate and ultimate analysis to produce valuable syngas. The results of different biomasses concerning proximate and ultimate studies conducted by different researchers are given in Table 4. The studies were conducted by various researchers for elemental analysis on MSW [31,32,33], sewage sludge [34,35,36], animal waste [37,38,39,40], wastewater [41,42], and agricultural waste [43,44,45,46], and the proximate values were investigated for MSW [31,32,33], sewage sludge [47,48], animal waste [40], wastewater [41], and agricultural waste [46].

2.2. Proximate and Ultimate Analysis

High ash-containing feedstock leads to difficulties during gasification including the plugging of the reactor, catalyst sintering, and appropriate ash disposal. Furthermore, by increasing the ash content of poultry manure from 17.2% to 25.1%, the gasification efficiency declined from 63% to 33%, and the concentration of H2 and CO decreased meaningfully. Furthermore, the decrease in the HHV of the syngas reduced from 4.3 MJ/m3 to 2.6 MJ/m3. The value of the volatile matter in biomass is another component taken into consideration. The increase in volatile matter results in rises in tar formation during the gasification of feedstock. The composition of tar primarily consists of aromatic compounds such as naphthalene, benzene, phenols, etc. The primarily aromatic compounds evaporate and then condense under gasification conditions. Moreover, shorter chain organics lead to the generation of tar at low temperatures such as aldehydes, ketones, and acids. Municipal solid waste 77–86 wt% and agricultural waste 73 to 83 wt% result in the production of a large amount of tar due to the higher value of volatile matter. During the gasification of biomass, fixed carbon controls the rate of gasification and the production of syngas [49]. Furthermore, volatile matter, fixed carbon, and ash content directly impact the gasification efficiency, gas yield, and inhibitory compound production. Therefore, the proximate analysis of the feedstocks is performed to thoroughly examine and select appropriate reaction conditions, reactor configurations, and catalysts.
During the gasification of feedstock, the higher concentration of oxygen and carbon tends to the higher amounts of CO and CO2 in syngas composition. While temperatures below 600 °C may cause CH4 production, the concentration of CH4 decreases, and CO and CO2 formation increase because of the methane decomposition and reforming reactions. Therefore, the gasification of municipal solid waste and agricultural residues produce more CO and CO2 as compared with other waste feedstocks. The greater amounts of sulfur and nitrogen in the feedstock tend to produce a greater amount of Nox and SOx poisoning in the catalyst. The presence of nitrogen is mostly in the form of organic complexes, and it reacts with hydrogen during gasification to generate hydrogen cyanide and ammonia. In addition, nitrogen is released in small quantities in the molecular nitrogen form such as aromatic organic compounds and nitrogen oxides, whereas a trivial quantity of N2 is retained as unreacted in the feedstock solids [49,50]. Under gasification conditions, sulfur is released in the form of H2S, which creates problems during the gas separation and treatment. In biomass feedstocks, the sulfur content is less than 1.5 weight %. Furthermore, animal waste products and sewage sludge contain sulfur ~1 wt. % and 0.5 wt. %. Higher amounts of sulfur lead to gas treatment issues and catalyst poisoning during the gasification of feedstock [50]. Maitlo et al. [51] investigated rice husk, cotton stalks, sugarcane bagasse, and saw dust. The results are given in Table 4.
Table 4. Shows the proximate analysis and ultimate analysis results. Data were taken from ref. [51,52].
Table 4. Shows the proximate analysis and ultimate analysis results. Data were taken from ref. [51,52].
BiomassProximate Analysis (%)Ultimate Analysis (%)
MoistureVMFCAshCHSNO
SD11.9861.3614.3612.3047.335.410.320.7946.15
RH7.1065.398.9618.5534.335.360.590.2959.43
SB5.9777.1912.304.5441.735.820.300.1052.05
CS5.5869.9816.318.1344.105.960.190.3649.39
VM = volatile matter; FC = fixed carbon; C = carbon; H = hydrogen; S = sulfur; N = nitrogen; O = oxygen; SD = sawdust; RH = rice husk; SB = sugarcane bagasse; CS = cotton stalks.

3. Gasification Agents

The process of biowaste gasification can be categorized as exothermic or endothermic. In addition, gasification may be classified by the gasifying agent, such as oxygen, steam or air, used during the gasification of biowaste. The application of different gasifying agents yields different syngas compositions, by-products, and heating values. The applications of different gasifying agents and their impact on syngas composition are summarized in Table 5 [31,53].

3.1. Gasifying Agent (Steam)

The data given in Table 5, comparing with the gasification of air and oxygen, shows that the amount of H2 produced by steam gasification is larger and the calorific value is higher [54]. This trend is caused by the role of steam in stimulating water–gas reactions, WGSR, methanation reaction, and steam reforming reactions. The primary reactions for example water–gas reactions and steam decomposition are the main factors that lead to the oxidation of feedstock. The change of organic carbon and hydrogen from the decomposition of steam into H2 and CO is primarily endorsed by the above reactions. Once the raw material is burned, the WGSR is liable for the transformation of gaseous H2O and CO into H2 and CO2. Garcia et al. [55] found that, as the S/B ratio was enhanced, the H2 and CO2 concentration in syngas composition was also increased, whereas the content of CO and CH4 in syngas composition reduced. Hernandez et al. also confirmed these trends [56] by gasifying the dealcoholized marc of grape. The study found that, as S/B increased, the H2/CO and H2/CO2 ratios increased, while the CH4/H2 ratio decreased. This change in reaction mechanism is the effect of the steam reforming of tars and the steam reforming of methane. During the gasification process, the reforming reactions of carbon and methane are intensified, resulting in a reduction in CH4 and a rise in H2 formation. Furthermore, the rise in S/B leads to a decrease in the CO/CO2 ratio, indicating that the WGSR has become dominant. In addition, steam gasification leads to a reduced content of recalcitrant by-products during the gasification of biowaste. Lapania et al. [57] observed that, when the S/B ratio was enhanced from 0.5 to 1.0, the tar formation lessened by 49%, and the char concentration reduced by 76%. This is because the steam initiates various condensation reactions with carbon, which decrease the carbon content [58]. Roche et al. [34] also found that steam reduces the amount of tar from 17% to 24%, compared to air. Hence, it was concluded that steam plays a vital role in making more hydrogen and methane during its gasification process.

3.2. Gasifying Agent (Air)

The most common gasifying agent is air, as it is copious and easy to use. Air-based gasification performance is highly dependent on temperature and equivalent ratio (ER). In particular, the higher the temperature of the air presented into the gasifier, the greater the heating value of the dry syngas produced [59]. Since air comprises up to 79% nitrogen, the generated gas is extremely diluted, which upsurges the treatment cost of gas separation [60]; thus, air gasification has serious flaws. Furthermore, the gaseous products formed by air gasification have a low calorific value of 3.50–7.80 MJ/m3. Hence, the use of air as a gasifying medium is generally limited to low heating and power production [61]. Mohammed et al. [62] observed that when the ER was enhanced from 0.150 to 0.350, for the gasification of empty fruit clusters, the production of tar decreased from 78% to 62% and for coal from 13% to 3%. Adding oxygen to the air will cause oxidization of the tar and its compounds, thus decreasing the tar and tar content during CO and CO2 production. Other research studies have confirmed the influence of ER on tar production [53,63,64]. With the increase in the ER, the gasification efficiency (GE) and the CCE incline to rise and then decrease.
Gao et al. [65] found that when ER was raised from 0.180 to 0.220, GE enhanced from 61.43% to 68.15%, then reduced to 59.56% as ER was raised to 0.280 in pinewood chip gasification. Zhao et al. [66] observed a similar inclination in the CCE value of wood chip gasification. When ER was enhanced from 0.220 to 0.310, CCE improved from 81% to 91.5%, but when ER increased to 0.34, CCE decreased to 88.5%. This trend of GE and CCE can be elucidated by suppressing the heat transfer between the solid particles. With the addition of more air to the gasifier, the inhibited heat transfer between solid particles and the air was observed which may result in a reduction in GE and CCE. In the process of air gasification, the H2 and CO concentration will first rise to the maximum value with the increase in ER, and then decline. With the increase in ER, the mole fractions of CH4 will continue to decrease, while the content of CO2 will continue to increase [63,67,68]. Zhao et al. [66] confirmed these trends, which showed that with an increase in ER from 0.22 to 0.34, the CO concentration reduced from 25.7 vol% to 21.5% vol, while the CH4 concentration in syngas lessened from 2.45 vol% to 0.87 vol%, and H2 concentration declined from 14.6 vol% to 10.2 vol%. At the same time, the CO2 concentration slowly improved from 11.70% to 12.30%, while LHV dropped from 6.67 MJ/m3 to 4.650 MJ/m3 with an increase in ER [66].

3.3. Gasifying Agent (Oxygen)

Oxygen is often used as a gasification agent, because it can produce syngas with a medium heating value. In this regard, the air-to-fuel ratio (AF) and ER are very important parameters to be calculated using Equations (17) and (18), respectively. The precise trend of increase in ER and its effects on gas characteristics under oxygen and air is identical. Niu et al. [53] observed that when the ER changes between 0.18 and 0.230, the calorific value of the generated gas can reach 8–10 MJ/m3. Moreover, an ER higher than 0.230 will cause the HHV of generated gas to drop sharply. The reason for this is that, with the increase in ER, the gasification reaction starts to favor the oxidation reactions, which will greatly increase the CO2 content in syngas composition [53]. Gao et al. [65] also confirmed the same and reported that H2 and CO content decreased, and CO2 content increased when ER was enhanced from 0.050 to 0.30. As compared with the gasification of air, the gasification of oxygen results in higher CCE. Nonetheless, two disadvantages of gasification using oxygen as a gasifying agent are: (1) the purchasing of pure oxygen is costly, affecting the overall operating cost of the plant, and (2) the separation of oxygen from synthesis gas is difficult and costly. Hence, oxygen is usually used with steam as a gasifying agent.
AF = m a m f = m a ˙ m f ˙
ER = = Air Fuel Stoichiometric Air Fuel Actual
where:
AF = Air to fuel ratio
ER =   = Equivalence ratio
ma = Mass of air
mf = Mass of fuel
m a ˙ = Mass flowrate of air
m f ˙ = Mass flowrate of fuel
From the above discussion, it has been established that air has serious disadvantages over the rest of the gasifying agents, due to the higher costs for the post-separation of high nitrogen. Oxygen alone is an expansive choice for the generation of pure oxygen and the separation of oxygen from syngas. Steam produces a good quantity of H2 and CH4 in syngas, and with less CO in syngas steam alone has few disadvantages. So, the most feasible gasifying agent is an appropriate mixture of oxygen and steam. This will improve the performance of the gasifier as well as enhance the overall economics of the process.

4. Catalyst Selection

Gasification reactions are exothermic, and therefore different types of catalysts are used for reducing the activation energy of reactions involved in the gasification process. The application of catalysts improves the syngas yield, CCE, and H2 content in syngas. Besides that, the incorporation of catalysts removes recalcitrant gasification by-products, including tar and char. The main types of catalyst are homogenous and heterogeneous catalysts. Homogeneous catalysts are those which occur in the same phase (gas or liquid) as the reactants, or have the same phase as the reaction mixture, while heterogeneous catalysts have a different phase. The selection of catalyst incorporation depends upon the preferred gaseous product [35]. The different types of catalyst, such as homogenous and heterogeneous, and their gasification performance are given in Table 6.

4.1. Heterogeneous Catalysts

Olivine (2MgO·SiO2), dolomite (MgCO3·CaCO3), and shells are natural mineral-based catalysts that are used directly or with some pretreatment techniques including calcination. Natural catalysts are abundantly available and cost-effective, and are widely used for tar removal from producer gas [69]. Moreover, the catalytic efficiency of natural catalysts can be increased through calcinating for 4 h at a temperature of 900 °C. The application of dolomite as a catalyst enhances H2 formation during the gasification of biomass.
Noble metal-based catalysts involve Pt, Rh, and Ru that possess high catalytic activity, having high sulfur resistance and high stability in the steam reforming of tar. Among the noble metal-based catalysts, Rh was observed as more significant compared with other catalysts. Noble catalysts are expensive compared to other catalysts.
Furthermore, besides noble catalysts, transition metal-based catalysts are also used for the reforming of tar. Transition metal-based catalysts include Cu, Co, and Fe. Transition metal-based catalysts are efficient in the steam reforming of tar, and are deactivated easily during carbon deposition in the case of high tar content in syngas.
The leading attraction of the above catalyst is that once the gasification reactions are quenched, the used catalyst can be recovered easily. Since heterogeneous natural catalysts are abundantly available, low cost, and can reduce tar content in the obtained syngas, they are most commonly used. Guan et al. [70] showed that in the application of dolomite to gasify municipal solid waste, the reduction in tar content was from 9.71 weight% to 0.0 weight%, thereby increasing the gas production from 0.620 to 1.470 m3/kg feedstock, while the H2 content increased from 30.6 to 50.2%. With the porous nature and higher content of alkaline components in MgO, Al2O3, CaO, etc., dolomite helps the cracking of tar compounds, which can participate in tar reduction. Heterogeneous catalysts are recognized to decompose tar to yield gaseous products.
Wu et al. [44], observed that, using a nickel catalyst, the H2 content touched 52% with a volume at 750 °C, reached 950 °C under non-catalytic gasification, and the tar content decreased from 27.30 g/m3 to 8.30 g/m3 in syngas with no catalyst in it. Although, it is well known that Ni deactivates and sinters during the reaction, so catalyst support is used to reduce fouling. Wang et al. [32] found that by using NiO/MD for MSW gasification, the syngas production was increased from 0.78 to 1.62 m3/kg MSW, and the H2 production was increased from 21.9 to 80.7 g H2/kg MSW. Compared with gasification using catalysts, tar production was increased from 38.7 to 0.23 g/m3 [32]. NiO/MD is not only conducive in the cracking of tar and the generation of H2, but also for alkaline oxides including MgO, CaO, etc., on dolomite support, which can prolong the service life of the catalyst without affecting the results when used for 10 h. Ni-Al2O3 is recognized as an effective catalyst for the decomposition of tar. The comparison of results when not using a Ni-Al2O3 catalyst in AWP gasification observed that H2 production improved from 3.90 to 16.40 mmol/g of manure, and CCE was improved from 23.5 to 40.6%. The comparison of gasification results under non-catalytic conditions showed that the tar transformation rate improved from 27.70 to 1.50 weight% [37].

4.2. Homogeneous Catalysts

The application of homogenous catalysts leads to higher mole fractions of H2 in producer gas composition. These catalysts are low-cost as compared with heterogeneous catalysts; however, homogenous catalysts are challenging to reclaim from reactions. Homogenous catalysts result in corrosive conditions. Compared with heterogeneous catalysts, homogenous catalysts are not frequently employed because of some drawbacks.
Hu et al. [71] used a CaO catalyst and inspected that when the ratio of the catalyst to biowaste improved from 0.0 to 0.70, the H2 mole fraction and gas production improved from 10.70 vol% to 64.70 mL/g MSW at 49.40 vol% and 277.70 mL/g MSW. When the ratio of catalyst to raw material exceeded 0.70, the H2 content began to decline. This increase can be described by the fact that a large amount of CaO makes it difficult to transfer heat. CaO is a tremendous catalyst for the selective separation of H2, because it is a CO2 adsorbent. In the steam gasification process of biological waste, NaOH is also used as a catalyst to promote H2 production. Gai et al. [36] performed gasification using SS and stated that NaOH stimulated the WGSR and improved the H2 mole fractions in syngas composition to 32.40 vol%, as compared to 24.0 vol% without a catalyst. Furthermore, it has been witnessed that the use of NaOH and other alkaline catalysts, such as KOH, reduces the mole fractions of CO and CO2 in syngas composition [36].
Table 6. Synthetic methods and performance of different catalysts.
Table 6. Synthetic methods and performance of different catalysts.
BiomassConditionsCatalystH2 YieldsMethodsRef
Wood sawdustT1 ¼ 535 °C, T2 ¼ 800oCNiZnAlOx48 vol%Co-precipitation[72]
Wood sawdustNiO loading: 7.2 wt%; 850 °CNiO/MgO51 vol%Commercial[73]
Corn stalkNi:Mg:Al ¼ 1:1:1; T1 ¼ 400 °C, T2 ¼ 800 °C; S/C ¼ 3.54; 30 minNi–Mg–Al56%Co-impregnation[74]
CorncobNi loading: 18.0%; 1 h; 650 °C; 1 g biomass; Steam: 30 kPaNi/Resin61 mmol/gIon exchange[75]
CorncobNi loading: 17.32%; 1 h; Steam: 30 kPa; 650 °C; 1 g biomassNi/lignite60 mmol/gChar Ion exchange[76]
CorncobNi loading: 6%; Ar atmosphere; 650 °C; 1 g biomassNi/dolomite22 mmol/gIon exchange[77]
Maize stalkNi, Ce loading: 14.9%, 2.0%; 900 °C, S/C ¼ 6; WHSV ¼ 12 h−1Ni–Ce/Al2O371%Co-impregnation[78]
Pine sawdustT ¼ 650 °C; GHSV ¼ 13000 h−1; S/C ¼ 7.6; Mg/Al ¼ 0.26; Co/Ni ¼ 0.10Ni/Co–Al–Mg-Co-precipitation[79]
Pine sawdustNi loading: 9.92%; 700 °C; S/C ¼ 12Ni/La2O3-αAl2O396%Impregnation[80]
Pine woodS/C ¼ 5.58; 650 °C; Ni loading: 28%Ni/Al catalysts77%Co-precipitation[81]
Pig manureNi loading: 19 ± 1 wt%; 650oC; ArNi/lignite69 mmol/gChar ion exchange[82]
SawdustS/C ¼ 5.0; WHSV ¼ 1.5 h−1; 800 °CNi/dolomite73%Impregnation[83]
Rice hullNi loading: 12%, Ce loading: 7.5%; W/B ¼ 4.9, 800 °CNi/CeO2–ZrO270%Impregnation[84]
Wood sawdustNi loading: 40 wt%, 0.25 g; 800oC; water: 5.0 mL/hNi/MCM-4151 vol%Impregnation[85]
SawdustS/CH4 ¼ 2; 800 °C; Catalyst: 15.0 g; GHSV ¼ 3600 h−1Ni/MgO81%Commercial[86]
The application of catalysts decreases the reaction temperature and enhances the carbon conversion rate; moreover, it increases the hydrogen selectivity from biomass for both steam gasification and supercritical water gasification. The addition of catalyst promotes tar cracking and the reforming of condensable fractions. Furthermore, the introduction of steam increases syngas production, particularly increasing hydrogen production. Therefore, catalytic steam reforming is a commonly used technique, as it removes tar effectively and simultaneously increases the selectivity of H2. Hence, the syngas contains lower CO and CH4 because of the steam reforming of methane and the water–gas shift of CO. Moreover, H2 formation increases during the catalytic gasification of biomass.

5. Types of Gasifiers

Many types of gasifiers are commonly used for the gasification of biomass through thermochemical conversion technologies. Depending upon the moisture content, ash content, shape, size, type of raw material, and requirements of the users, gasifiers are mainly divided into two categories, i.e., fixed bed and FBG [87,88]. A few other types of gasifiers are also employed for the gasification of biomass, including rotary kiln, plasma, and EFG [89,90]. Brief schematic diagrams of biomass gasifiers are given in Figure 3, keeping in view the advantages and disadvantages of reactors for the catalytic gasification of biomass. Biomass gasification is performed in various gasifier configurations such as FBG, fixed bed, EFG, and plasma gasifiers. EFG is widely engaged in biomass gasification due to good mixing capabilities, higher heating and mass transfer rates, scalability, feed flexibility, high carbon conversion efficiency, and high reaction rates.

5.1. Fixed-Bed Gasifiers

Fixed-bed gasifiers are further divided into three categories, i.e., updraft, downdraft, and cross draft gasifiers. These categories of fixed-bed gasifiers are because of the interaction between biomass and gasifying agent [93].

5.1.1. Updraft Gasifier

The oldest and simplest gasifier used for synthesis gas production is an updraft gasifier. A gasifying agent is introduced from the inlet point located at the bottom of the gasifier and moves upward in an updraft reactor, whereas the raw feedstock is introduced from the top of the gasifier and travels downward through gravity [94]. Thus, gasifying agents and biomass create a counter-current flow within the gasifier. Biomass and gasifying agents pass through all the zones of the gasifier, such as the drying, pyrolysis, reduction, and oxidation zones, and different types of reactions occur in all zones of the updraft gasifier. The main strength of the updraft gasifier includes a low-pressure drop, high thermal efficiency, and low slag formation. At high flame temperature, the updraft gasifier tends to produce low dust content in syngas. The major drawbacks of updraft gasifiers are long engine startup times, low syngas production, and high tar sensitivity [95]. The schematic diagram of the updraft gasifier is shown in Figure 3a.

5.1.2. Downdraft Gasifier

The downdraft reactor is commonly called a concurrent gasifier in which the feedstock and reacting gas passage together downward in the same direction. Comprising gas and tar, all decomposition products move through the oxidation zone of the downdraft gasifier, and the cracking of feedstock occurs continually. In a downdraft reactor, high-quality synthesis gas with low tar content can be achieved compared with an updraft gasifier. The fixed-bed downdraft reactor is commonly used for small-scale power generation [96,97]. The key constituents of biomass (lignin, hemicellulose, and cellulose) decompose, producing H2, CO2, CH4, and H2O by the reaction of hydroxyl and methoxy. The schematic diagram of the downdraft gasifier is shown in Figure 3b.

5.1.3. Cross Draft Gasifier

Compared to “updraft” and “downdraft” fixed-bed reactors, the cross draft reactor is the simplest type of feedstock gasification reactor. The feedstock is introduced from top to bottom and the thermochemical transformation of the biomass takes place continuously within the reactor. As shown in Figure 3c, the downdraft reactor is configured to have an ash zone, combustion, and a reduction section. In comparison with the “updraft” and “downdraft” types, the air is injected into the descending reactor from the side. The cross draft reactor has excellent flexibility in synthesis gas production. However, biomass fuel content is high, volume is small, and gasification is difficult [98].

5.2. Fluidized Bed Gasifiers

5.2.1. Circulating and Bubbling Fluidized Bed Gasifiers

The bubbling fluidized bed gasifier is kept fluidized by inert materials such as dolomite, olivine, sand, SiO2, etc. The agent used for gasification is located from the bottom of the bed through the distribution network at varying flow rates. The inner bed is kept fluidized to achieve an appropriate reaction with raw material. The inert liquid-like material is constantly agitated in the existence of gas bubbles to provide uniform heat transfer conditions for gas and biomass. The circulating bed gasifier at this point is alike to the BFBG, as shown in Figure 3d,e. The separator cyclone is installed at the reactor outlet to bring the biomass to the bubbling reactor. The high-velocity gas energy, usually 5–10 m/s, can be provided from combustion to drive the pyrolysis and gasification of the biomass. Presently, the FBG is a capable reactor for achieving tar-free syngas [99,100]. It is characterized by feed flexibility, higher mass and heat transfer efficiency, a uniform temperature within the gasifier, and high syngas yield. Biomass such as rice straw, grass, and wheat straw, having a high content of ash, can damage equipment and reduce the quality of syngas production [101]. The temperature at which FBG is operated is usually maintained from 800 °C–900 °C, which is determined according to the melting point. Since the gasification reaction usually takes place at high temperatures, high activity catalysts must be planned during the gasification of feedstock.

5.2.2. Entrained Flow Reactor

Feedstock with a size less than 1 mm along with gasifying agents are fed in co-current EFG. The usual operating pressure for EFG is controlled from 25 to 30 bar and the temperature of the reactor is normally maintained from 1300 °C to 1500 °C. The size of the feedstock is reduced to make its slurry easy, then the slurry is introduced into the EFG. More specifically, EFG is employed for the treatment of feedstocks including biomass, coal, mixed plastic waste, forest residues, sewage sludge, and agricultural residues for syngas production. The removal of moisture through heat treatment is necessary before the application of selected raw material for efficient gasification [102,103]. Figure 3f shows the arrangement of the gasifier for biomass gasification.

5.2.3. Plasma Reactor

In a plasma gasifier, two electrodes are commonly installed such as carbon electrodes or Cu electrodes. Furthermore, when the internal temperature of the reactor approaches up to 10,000 °C, the plasma gasifier will provide an electric charge commonly known as “arc”. The atomic degradation of feedstock takes place in gasifiers during gasification under different atmospheres. In the absence of the oxidizing agent, the energy for biomass temperature rise is provided through the plasma process. In the following two situations, the plasma reactor for the gasification of biomass may be applied. (1) Plasma gasifiers can make gaseous products clean during the upstream process of biomass gasification. Moreover, tar degradation occurs in light components. (2) The temperature of the gasifier can be controlled at varying moisture content, flowrate, size of feedstock, gasifying agent, and syngas quality, independently. The plasma reactor is shown in Figure 3g. The reactor can treat general, medical, and hazardous waste, and can achieve full carbon conversion regardless of the type of raw material. This plasma technology can reach very high temperatures and decompose raw materials completely into CO and H2, producing clean synthesis gas [104]. The use of higher plasma power rises the H2/CO ratio, resulting in higher temperatures and greater cracking of the tar. Under different operating conditions, the HHV of syngas produced is between 6 and 7 MJ/Nm3. It has been determined that the H2/CO ratio produced is much higher than that of conventional reactors using air, oxygen, and steam as gasifying agents [105]. During the gasification of hazardous industrial waste under plasma gasification conditions, the H2/CO ratio is observed to be approximately 1.0. A study conducted by Mountouris et al. [106] found that the use of sewage sludge under plasma gasification conditions had a similar gaseous composition.

5.2.4. Rotary Kiln

The rotary kiln gasifier is an important type of reactor, and it is extensively used on industrial scale waste gasification. The rotary furnace gasifier is usually a steel shell having a wear-resistant refractory structure that can decrease the temperature of the metal shell. As shown in Figure 3h, dry and fixed-size feedstock materials are organized and fed by maintaining an appropriate feeding rate. The generated gas and slag are commonly discharged from the top and bottom of the gasifier outlet. Due to the continuous rotation of the reactor, exposure to the surface of the biomass and gasification agent can be fully contacted. Nevertheless, the flue gases formed by gasification can take away a lot of heat. Therefore, unless a barrier is installed in the reactor, the heat interchange between biomass and reaction gas is ineffective [87].
Fixed-bed gasifiers, such as updraft and downdraft gasifiers, are less efficient during gasification in terms of carbon conversion efficiency, cold gas efficiency, and higher heating value of syngas produced. Furthermore, fluidized bed gasifiers are efficient and produce high-quality syngas. Particularly, entrained flow gasifiers are most efficient in terms of operation, H2 production, and product purity.

6. Performance of a Gasifier

The performance of a gasifier is evaluated through the consideration of various parameters such as cold gas efficiency, CCE, gas production, HHV, LHV, exergy efficiency, and equivalence ratio. Syngas production (GV, m3/kg) is the ratio between the volume of syngas under standard condition (Vg, m3) and the feeding of biomass into the gasifier (Mbiomass, kg). The gas production is further demonstrated in Equation (19).
G v = V g M biomass  
The syngas lower heating value is the chemical energy contained in the tar-free cold syngas. The total calorific value of syngas is given in Equation (20). It is the sum of the LHV of syngas individual components (LHVi) multiplied by the vol.% of each component gas i (xi):
LHV syngas = Σ   x i ×   LHV i
while (HHVsyngas) is the sum of the LHV of syngas and the latent heat of water vapor formed during combustion of feedstock.
ER is equal to the amount of oxygen introduced into the reactor in comparison to the amount of oxygen required for the complete combustion of feedstock. The generic formula of biomass is assumed to be CHxOy, for defining the stoichiometric reaction of biomass with oxygen. Equation (21) is the stoichiometric reaction of biomass from which the amount of oxygen is calculated with a fixed amount of fuel. Then, the calculated values are used in Equations (17) and (18) for the estimation of equivalence ratios (ER).
CH X O y + 1 + x 2 y 4   O 2     CO 2 + x 2   H 2
In the gasification process, the CCE is the % of carbon in feedstock converted into syngas. Equation (22) represents carbon conversion efficiency [20,48].
CCE = Carbon   in   syngas   ×   syngas   flow   rate carbon   content   in   biomass   ×   biomass   flow   rate × 100  
In the gasification process, CGE is the ratio between the chemical energy in syngas and the chemical energy contained within the feedstock, and is demonstrated in Equation (23).
CGE = LHV syngas   ×   syngas   flow   rate LHV biomass ×   biomass   flow   rate × 100  
The exergy efficiency is the last parameter calculated during the efficiency measurement of the gasifier. The exergy is the estimation of the actual potential of the system to perform work. It is the ratio of energy that flows out and into a system. In exergy efficiencies, both physical and chemical exergy contributions are taken into consideration at high operational temperatures. For the adiabatic gasifier, Equation (24) demonstrates the exergy efficiency.
ψ = n syngas   e ch ,   syngas   + e ph ,   syngas M biomass × e ch ,   biomass   +   n agent × e agent  
In Equation (6), n syngas   is syngas molar amount (kmol), e ch ,   syngas   + shows the chemical exergy of syngas (kJ/kmol), M biomass represents the feeding of biomass (kg), e ph ,   syngas shows the physical exergy of syngas (kJ/kmol), e ch ,   biomass   is biomass chemical exergy (kJ/kg), n agent is the gasifying agent molar amount (kmol), and e agent shows gasifying agent chemical exergy (kJ/kmol).

7. Effect of Operating Parameters on Syngas Composition

7.1. Effect of Temperature

The temperature of gasification is one of the most significant parameters that affect the quality of the gas produced and the efficiency of the process. As most of the gasification reactions are endothermic, the temperature rise will increase the endothermic cracking reaction of tar, and result in a reduction in the tar content producing H2, CO, and CO2 in synthesis gas composition. Methane decomposition reactions are highly endothermic, so CH4 can be maintained at lower temperatures, i.e., 300–600 °C. Furthermore, when the temperature of the reactor is maintained higher than 600 °C, CH4 will be decomposed through the reforming reaction of methane and hydrocarbons [107,108]. On the other hand, H2 increases with the increase in overall temperature, mainly due to the steam reforming, hydrocarbon reforming, and methane decomposition reactions. Although the production trends of H2 and CH4 are well represented in the literature, there is still no clear trend for CO and CO2. The effect of temperature on the non-catalytic steam gasification of organic waste was investigated by different researchers [44,109,110,111]. With the increase in gasification temperature, all biowaste materials follow the theoretical trend of H2 and CH4. For SS gasification (700–1000 °C), H2 increased by 7.0% and CH4 decreased by 4.0% [109]; H2 of SS and pine chips (600–1000 °C) increased by 11.0%, and CH4 decreased by 2.5% (900 °C) [110]; for wood waste (750–950 °C), H2 increased by 19.7% and CH4 decreased by 5.1% [51]; legume straw H2 increased by 15.0% and CH4 decreased by 4.0% (750–850 °C) [111]. Though, as anticipated, there was no clear trend in the composition of CO2 and CO. With the gasification using steam as a gasifying agent for the mixture of SS and pine sawdust, and SS alone, only SS displayed a decrease in CO2 and an increase in CO within the temperature range, but the gasification of wood residues and legume straws presented an increase and decrease in CO2. Some studies have shown that, as CO content increases, temperature also increases, which reduces CO2 content [112,113], while others have shown a contrary tendency [114]. The upsurge in CO at the cost of CO2 is endorsed to the Boudouard reaction; though, because this reaction directly competes with the WGSR, which is supposed to be dominant above 750 °C, these trends remain ambiguous.

7.2. Effect of Pressure

Considering the downstream use of the produced gas, biomass gasification is usually carried out under high atmospheric pressure. Some applications of product gas are performed through Fischer–Tropsch synthesis to convert gas into methanol or synthetic diesel, requiring high-pressure syngas production where the gasification of biomass under atmospheric pressure is advantageous. Moreover, the rise in gasifier operating pressure decreases the production of tar in the produced gas. Furthermore, some studies performed in an FBG show that the content of tar mostly increases naphthalene as the pressure of the gasifier increases from 0.1 MPa to 0.5 MPa, so the content of CO decreases, whereas the increase in the concentration of CH4 and CO2 was observed. A gasification model coupled with SOFC and gas turbines was carried out to prove that moderate pressures, for example, up to 4 bar, will not have a significant effect on the process of gasification. Interestingly, this affected the efficiency of the turbine, so the overall efficiency of the unit improved from 23% to 35% [115].

7.3. Effect S/B Ratio

Among various operating parameters for syngas yield, the S/B ratio is the most significant operating parameter during the gasification of biowaste. Moreover, the main reaction for H2 production is the steam reforming reaction. Other major reactions that are involved in making H2 include: cracking of tar, water–gas shift reaction, and water–gas reactions. The effect of the S/B ratio on syngas composition was investigated by Guan et al. and Tursun et al. [70,116]. When the S/B ratio was enhanced, it resulted in a decrease in the mole fractions of CH4 and CO in syngas composition, while the rise in mole fraction of CO2 was increased. Guo et al. [117] stated, regarding the S/B ratio, that from 0.4 to 2.0, the mole fractions of CO lessened from 36.5 to 22.5%, the mole fractions of CH4 reduced from 2, 4% to 0.6%, and the mole fractions of CO2 improved from 13% to 20%. However, the H2 inclination is not monoatomic, as with other gas yield tendencies. The H2 mole fractions first increase as the S/B ratio increases and then drop after reaching a certain maximum H2 yield. Chen et al. [107] observed that when the acid hydrolyzed residue of corncob was gasified with steam when the S/B ratio increased from 1.3 to 5.3, H2 concentration increased from 38.4% to 45.6%. By further enhancing the S/B ratio ranging from 5.3 to 7.5, the H2 concentration dropped from 45.6% to 42.5%. The impact of the S/B ratio on tar yield, syngas production, CCE, and HHV was estimated by Guan et al. and Pfeifer et al. [70,118]. There is a critical S/B ratio that depends on the reactor configuration and application of the catalyst.
Many research investigations have shown that, by enhancing the S/B ratio greater than 1.0, the calorific value of produced gas is considerably reduced, because of the sharp reduction in the mole fractions of CO and CH4 [119]. In addition, previously conducted research studies have observed that CCE [70] and gas production rate [70] will first rise and then decline after the critical S/B value. In addition, tar production will decline and then rise after this value [99]. Consequently, knowing the critical value of S/B for particular gasification is essential for optimizing gas production. The increase in the S/B ratio reduces tar formation, and makes tar more phenolic, with more C–O–C bonds, and easier catalytic conversion [120]. Though, some other research studies disagree with this trend. Fuentes-Cano et al. [121] observed that steam cannot decompose light hydrocarbons and aromatic hydrocarbons, and has merely a trivial effect on the disintegration of non-aromatic tar compounds, which is consistent with earlier studies [122,123]. Fuentes-Cano et al. [121] found that when steam was added, the yield of gravimetric tar dropped from about 285 g/kg SS to about 250 g/kg SS, and from 110 g/kg at 900 °C. Commonly, the molar ratio of H2/CO produced by gasification is nearly 0.7. The effect of S/B ratio on H2/CO ratio of the syngas for biomass feedstocks was inspected by Ge et al. [45]. The type of biomass and S/B value have a substantial effect on the H2/CO composition of synthesis gas. Between S/B ratios of 1 and 2, an extensive variety of H2/CO ratios were witnessed. This may be caused by numerous aspects. For instance, different feedstocks, catalysts, and reaction environments will produce diverse H2/CO ratios, resulting in great variability. Within this S/B range, syngas may be used in diverse ways to obtain various types of end products, including diesel, gasoline, and methanol, through the FT process. When the S/B ratio is between 2 and 3, the syngas will have an H2/CO ratio more than 2, thereby producing high-value synthetic gaseous products H2 and ammonia, etc. This means that when S/B ratio (>2) rises, the role of the catalyst and reaction conditions in determining the H2/CO ratio will weaken. Figure 4 shows the effect of S/B ratio on syngas composition during the gasification of biomass. While Figure 5 shows the effect of S/B ratio on syngas production, LHV, CCE, and tar production for biowaste gasification.

7.4. Energy and Exergy Efficiency

Exergy and energy efficacy are the two main performance gauges that are commonly used to assess dissimilar processes in the chemical engineering, transportation, and agricultural sectors. Energy is the ability of a system to perform work, while exergy is the maximum valuable work that can be achieved when the scheme changes from one state to equilibrium. The exergy energy in the reaction process is the total energy in the synthesis gas divided by the sum of the energy in raw material and processing energy. S/B is a parameter that strongly affects exergy efficiency. Husseini et al. [125] reported that in sawdust gasification when S/B was enhanced from 0.15 to 0.20, the exergy efficiency dropped from 28.0% to 17.0%. However, further research has observed that there is an optimal S/B value that can maximize energy efficiency. Figure 6 shows the effect of S/B on the exergy efficiency of biowaste using a mixture of wood pellets (1000 °C) and eighty-six varieties of biomass at 700 °C [126]. Wu et al. [126] found that, at 1000 °C, when S/B reaches 0.53, energy and exergy value increase to about 83% and 75%, respectively. Although, when S/B increases, both values drop to 71% when S/B is enhanced to 2.50. Heat recovery is another parameter that affects exergy efficiency.
Song et al. [127] observed that the exergy efficiency, devoid of heat reclamation, extended the highest value of 52.1, 49.9, 48.5, and 46.4% at 700, 750, 800, and 850 °C, and at these temperatures, the heat recovery rate of steam reached 65.7, 63.9, 63.3, and 61.7% respectively. Compared with the reaction without heat recovery, heat recovery resulted in a 14% escalation in exergy efficiency. The recovery of heat has significant consideration for making most of the energy efficiency and exergy of the gasification process. The suitable S/B value must be determined experimentally to maximize the energy exergy/efficiency, H2 selectivity, HV, and the resulting syngas yield.

7.5. Formation of Tar and Removal Techniques

Producer gas (PG) tar is very common in the biowaste gasification process, and it is considered to be the main technical obstacle for the expansion and application of industrial-scale gasification technology [24,128,129,130]. Tar is a combination of highly aromatic condensable organic compounds which are made during the partial oxidation of feedstock [120]. The composition of tar largely depends on the temperature of the reactor and is mainly divided into primary products such as phenol, secondary products such as benzene, toluene, and xylene, and tertiary products such as pyrene, indene, and naphthalene at 200–500 °C, 500–1000 °C, and over 700 °C, respectively [120,131]. Primary tar and tertiary tar are commonly exclusive, and before tertiary products appear the primary products are destroyed [120]. While the PG temperature is less than 400 °C, the tar condenses downstream of the reactor, causing various complications including catalyst deactivation, blockage, fouling, corrosion, and general failures of equipment [130]. Therefore, it is technically impractical to use crude PG with high tar content in certain applications, such as fuel and chemical catalytic synthesis, before the main refining steps. Condensation of tar in pipes and other process equipment also reduces the efficiency of the plant and increases operating costs. In addition, these composites can develop into intricate molecular activities through polymerization, which increases the exertion of removing them. Tar compounds comprise a substantial portion of energy, which is lost as they are removed from PG [132]. Therefore, the tar concentration in PG should be decreased to a value well-suited with the required downstream application. This may necessitate costly gas conditioning equipment, which makes the process economically unappealing. Furthermore, for direct application such as fuel in the combustion process of furnaces, tar is not a problem if condensation is evaded in the transportation pipeline that feeds the burner. Nevertheless, more innovative applications, for instance in internal combustion engines or secondary fuel synthesis, have strict limits on the tar concentration in PG. Therefore, for biomass gasification technology to take a profitable step forward and be integrated into different biorefining projects, it is imperative to improve cost-effective processes to eliminate or reduce tar formation. Besides that, tar compounds can also be altered into lighter gaseous substances including CO, CH4, and H2, meeting strict tar concentration limits. Various tar removal processes and technologies are being studied [133,134].
Mechanical treatments can achieve high tar elimination efficiency, but can considerably decrease the energy efficiency of a plant, generate hazardous waste, lessen gas production, and reduce profitability [135]. Nevertheless, even major processes are not adequate to attain complete tar elimination [136]. Furthermore, tar removal reactions are considered kinetically limited reactions. Hence, the rate of reaction can be increased by enhancing the temperature or by the application of catalysts. The catalytic decomposition of tar is a complex mix of equilibrium reactions and hydrocarbon decomposition. When the molecular weight of tar increases, the dew point also increases causing serious operating problems. Dry reforming reactions are catalyzed by metal catalysts. Furthermore, dry reforming is more favored as compared with steam reforming reactions above 830 °C. Moreover, catalytic filters have recently been developed for the elimination of tar from producer gas. In this method, the combination of filtration for particle elimination and catalytic tar cracking from producer gas is achieved in one step. The experimental results revealed that this method is efficient for the catalytic elimination of tar and particles from producer gas. Different operating parameters for gasifiers and their effects on syngas production are summarized in Table 7.

7.6. Carbon Conversion and Cold Gas Efficiency

Kuo et al. [97] inspected the gasification performance of three biomasses, including raw bamboo, torrefied bamboo at 250 °C, and torrefied bamboo at 300 °C using a downdraft gasifier, which was assessed by thermodynamic investigation. Two factors—modified Equivalence Ratio (ERm) and Steam Supply Ratio (SSR)—were considered to regarding their impact on biomass gasification. CGE, CCE, and HHV were used as indicators to detect gasification performance. Analysis showed that torrefied bamboo was beneficial to increase the production of syngas. The greater the torrefied temperature, the greater the production of syngas, but the lower the ERm value of TB at 300 °C. As the high calorific value of TB at 300 °C was considerably improved over that of raw bamboo and TB at 250 °C, the former possesses the minimum CGE amongst the three fuels. In the research scope of ERm and SSR, the CC value of raw bamboo and TB at 250 °C was always greater than 90%, but with the increase in ERm, more CO2 was produced, thereby reducing CGE. The maximum raw bamboo synthesis gas yield value and CGE, with TB at 250 °C and TB at 300 °C, were found for ERm: SSR was (0.2, 0.9), (0.22, 0.9) and (0, 28, 0.9) respectively. After considering the yields of syngas, CGE, and CC at the same time, predictions showed that TB at 250 °C was a more feasible gasification fuel.
Teh et al. [159] discussed that biomass gasifiers should be operated above 900 °C for the efficient conversion of carbon into syngas. Hoque et al. [160] achieved cold gas efficiency >60% for rice husk, and >70% for both sawdust and coconut shells. Konda et al. [161] worked on oil palm fronds, and obtained carbon conversion efficiency of 95% and cold gas efficiency of 59.1%, respectively. Figure 7 shows CGE, CCE, and LHV of (a) raw bamboo, (b) torrefied bamboo at 250 °C, and (c) torrefied bamboo at different temperatures.

8. Conclusions

The following conclusions can be drawn from the current review:
  • It is established that an appropriate mixture of pure oxygen with steam has advantages over other options for biomass gasification in terms of technical and economical perspectives;
  • The chemical composition of the biomass, reactor design, gasifier temperature, and catalyst work together to selectively separate H2 and improve gasification efficiency;
  • The commercial application of biomass gasification is presently restricted to municipal solid waste and agricultural waste;
  • The main problems facing the commercial prospects of bio-waste gasification include large amounts of tar in the gas produced, difficulty in separating individual gaseous compounds, and the deactivation of gasification catalysts due to compounds containing nitrogen and sulfur;
  • There are several challenges related to plasma gasification, and attention should be paid to the success and future commercialization of plasma gasification;
  • The catalytic gasification of biomass is a key technology for environmental and chemical production.
Future sustainable biomass availability is one of the key reservations when considering large-scale biomass usage. However, the increased demand for limited biomass resources raises the requirement for monitoring biomass supply. Policies to encourage the further replacement of fossil fuels with biomass in the industry should be designed. The following are important research criteria for further biomass usage on an industrial scale:
  • Develop and demonstrate advanced technologies to transform biomass into fuels and energy;
  • Pursue and develop techniques to further incorporate the use of biomass where it is technically viable;
  • Develop and optimize sustainable and cost-effective biomass supply;
  • Design market-based solutions to foster investments in biomass.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/su14052596/s1. Figure S1: Detailed review on biomass gasification. References [162,163,164,165,166,167,168,169,170,171,172,173,174,175,176,177,178,179,180] are cited in the Supplementary.

Author Contributions

G.M. performed data entry and manuscript arrangement; I.A. worked on the relevance of data regarding syngas composition; K.H.M. worked on different types of the gasifier and their performance evaluation; H.A.M. reviewed and performed the write-up of catalysts and their application; I.N.U. estimated and collected data regarding CCE, CGE, and HHV of syngas; S.A. and A.M.P. compiled and critically analyzed the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This study was conducted at Dawood University of Engineering and Technology, hence all authors are highly indebted for providing the working environment and facilities required for the successful completion of this work. Additionaly, Maitlo, H.A acknowledges support made by a grant from the National High-Level Foreign Experts Project (No. QN20200002003).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

H2Hydrogen MWMegawatt
COCarbon monoxideTBTorrefied bamboo
LHVLower heating valueSSRSteam supply ratio
oCCentigrade S/BSteam biomass ratio
CO2Carbon dioxideEREquivalence ratio
CH4Methane GEGasification efficiency
%Percent CCECarbon conversion efficiency
EFGEntrained flow gasifierVol%Volume percent
FBGFluidized bed gasifier HCHeterogenous catalyst
PGProducer gasNi Nickle
MJ/kg Megajoule per kilogram NiONickle oxide
MJ/LMegajoule per liter MgOMagnesium oxide
MJ/m3Megajoule per cubic meterAl2O3Aluminum oxide
US$/kg United states dollar per kilogramCaOCalcium oxide
MSWMunicipal solid waste NaOHSodium hydroxide
wt%Weight percentKOHPotassium hydroxide
HHVHigher heating valueSiO2Silicon oxide
NOxNitrogen oxideBFBGBubbling fluidized bed gasifier
SOxSulfur oxideCGECold gas efficiency
N2Nitrogen SSSewage sludge
H2SHydrogen sulfide MPaMegapascal
MJ/kmol Megajoule per kilomoleSOFCSolid oxide fuel cell
WGSRWater gas shift reactionCFBG Circulating fluidized bed gasifier
DDDowndraft UDUpdraft

References

  1. Furlan, F.F.; Filho, R.T.; Pinto, F.H.P.B.; Costa, C.B.B.; Cruz, A.J.G.; Giordano, R.L.C.; Giordano, R.C. Bioelectricity versus bioethanol from sugarcane bagasse: Is it worth being flexible? Biotechnol. Biofuels 2013, 6, 142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Dantas, G.A.; Legey, L.F.L.; Mazzone, A. Energy from sugarcane bagasse in Brazil: An assessment of the productivity and cost of different technological routes. Renew. Sustain. Energy Rev. 2013, 21, 356–364. [Google Scholar] [CrossRef]
  3. Seabra, J.E.A.; Macedo, I.C. Comparative analysis for power generation and ethanol production from sugarcane residual biomass in Brazil. Energy Policy 2011, 39, 421–428. [Google Scholar] [CrossRef]
  4. Akay, G.; Jordan, C.A. Gasification of Fuel Cane Bagasse in a Downdraft Gasifier: Influence of Lignocellulosic Composition and Fuel Particle Size on Syngas Composition and Yield. Energy Fuels 2011, 25, 2274–2283. [Google Scholar] [CrossRef]
  5. Ahmed, I.I.; Gupta, A.K. Sugarcane bagasse gasification: Global reaction mechanism of syngas evolution. Appl. Energy 2012, 91, 75–81. [Google Scholar] [CrossRef]
  6. IRENA; International Renewable Energy Agency. Renewable Energy Target Setting; International Renewable Energy Agency: Abu Dhabi, United Arab Emirates, 2015. [Google Scholar]
  7. Demirbas, A. Emission Characteristics of Gasohol and Diesohol. Energy Sources Part A Recovery Util. Environ. Eff. 2009, 31, 1099–1104. [Google Scholar] [CrossRef]
  8. Foust, T.D.; Aden, A.; Dutta, A.; Phillips, S. An economic and environmental comparison of a biochemical and a thermochemical lignocellulosic ethanol conversion processes. Cellulose 2009, 16, 547–565. [Google Scholar] [CrossRef]
  9. Demirbas, A. Biofuels sources, biofuel policy, biofuel economy and global biofuel projections. Energy Convers. Manag. 2008, 49, 2106–2116. [Google Scholar] [CrossRef]
  10. De Filippis, P.; Borgianni, C.; Paolucci, M.; Pochetti, F. Gasification process of Cuban bagasse in a two-stage reactor. Biomass Bioenergy 2004, 27, 247–252. [Google Scholar] [CrossRef]
  11. Demirbas, A. (Ed.) Thermochemical Conversion Processes. In Biofuels: Securing the Planet’s Future Energy Needs; Springer London: London, UK, 2009; pp. 261–304. [Google Scholar]
  12. Pinto, F.; André, R.N.; Carolino, C.; Miranda, M. Hot treatment and upgrading of syngas obtained by co-gasification of coal and wastes. Fuel Process. Technol. 2014, 126, 19–29. [Google Scholar] [CrossRef] [Green Version]
  13. Anukam, A.; Mamphweli, S.; Reddy, P.; Meyer, E.; Okoh, O. Pre-processing of sugarcane bagasse for gasification in a downdraft biomass gasifier system: A comprehensive review. Renew. Sustain. Energy Rev. 2016, 66, 775–801. [Google Scholar] [CrossRef] [Green Version]
  14. Mohanty, P.; Pant, K.K.; Naik, S.N.; Parikh, J.; Hornung, A.; Sahu, J.N. Synthesis of green fuels from biogenic waste through thermochemical route—The role of heterogeneous catalyst: A review. Renew. Sustain. Energy Rev. 2014, 38, 131–153. [Google Scholar] [CrossRef]
  15. Samiran, N.A.; Jaafar, M.N.M.; Ng, J.-H.; Lam, S.S.; Chong, C.T. Progress in biomass gasification technique—With focus on Malaysian palm biomass for syngas production. Renew. Sustain. Energy Rev. 2016, 62, 1047–1062. [Google Scholar] [CrossRef]
  16. Bocci, E.; Sisinni, M.; Moneti, M.; Vecchione, L.; Di Carlo, A.; Villarini, M. State of Art of Small Scale Biomass Gasification Power Systems: A Review of the Different Typologies. Energy Procedia 2014, 45, 247–256. [Google Scholar] [CrossRef] [Green Version]
  17. Sansaniwal, S.K.; Pal, K.; Rosen, M.A.; Tyagi, S.K. Recent advances in the development of biomass gasification technology: A comprehensive review. Renew. Sustain. Energy Rev. 2017, 72, 363–384. [Google Scholar] [CrossRef]
  18. Ahrenfeldt, J.; Thomsen, T.P.; Henriksen, U.; Clausen, L.R. Biomass gasification cogeneration—A review of state of the art technology and near future perspectives. Appl. Therm. Eng. 2013, 50, 1407–1417. [Google Scholar] [CrossRef] [Green Version]
  19. Puig-Arnavat, M.; Bruno, J.C.; Coronas, A. Modeling of trigeneration configurations based on biomass gasification and comparison of performance. Appl. Energy 2014, 114, 845–856. [Google Scholar] [CrossRef]
  20. Puig-Arnavat, M.; Bruno, J.C.; Coronas, A. Review and analysis of biomass gasification models. Renew. Sustain. Energy Rev. 2010, 14, 2841–2851. [Google Scholar] [CrossRef]
  21. Demirbaş, A. Biomass resource facilities and biomass conversion processing for fuels and chemicals. Energy Convers. Manag. 2001, 42, 1357–1378. [Google Scholar] [CrossRef]
  22. Kirkels, A.F.; Verbong, G.P.J. Biomass gasification: Still promising? A 30-year global overview. Renew. Sustain. Energy Rev. 2011, 15, 471–481. [Google Scholar] [CrossRef]
  23. Motta, I.L.; Miranda, N.T.; Maciel Filho, R.; Maciel, M.R.W. Biomass gasification in fluidized beds: A review of biomass moisture content and operating pressure effects. Renew. Sustain. Energy Rev. 2018, 94, 998–1023. [Google Scholar] [CrossRef]
  24. Ruiz, J.A.; Juárez, M.C.; Morales, M.P.; Muñoz, P.; Mendívil, M.A. Biomass gasification for electricity generation: Review of current technology barriers. Renew. Sustain. Energy Rev. 2013, 18, 174–183. [Google Scholar] [CrossRef]
  25. Phitsuwan, P.; Sakka, K.; Ratanakhanokchai, K. Improvement of lignocellulosic biomass in planta: A review of feedstocks, biomass recalcitrance, and strategic manipulation of ideal plants designed for ethanol production and processability. Biomass Bioenergy 2013, 58, 390–405. [Google Scholar] [CrossRef]
  26. Palma, C.F. Modelling of tar formation and evolution for biomass gasification: A review. Appl. Energy 2013, 111, 129–141. [Google Scholar] [CrossRef]
  27. Devendra, L.P.; Kiran Kumar, M.; Pandey, A. Evaluation of hydrotropic pretreatment on lignocellulosic biomass. Bioresour. Technol. 2016, 213, 350–358. [Google Scholar] [CrossRef] [PubMed]
  28. Sun, S.; Sun, S.; Cao, X.; Sun, R. The role of pretreatment in improving the enzymatic hydrolysis of lignocellulosic materials. Bioresour. Technol. 2016, 199, 49–58. [Google Scholar] [CrossRef] [PubMed]
  29. Kalinci, Y.; Hepbasli, A.; Dincer, I. Comparative exergetic performance analysis of hydrogen production from oil palm wastes and some other biomasses. Int. J. Hydrog. Energy 2011, 36, 11399–11407. [Google Scholar] [CrossRef]
  30. Faba, L.; Díaz, E.; Ordóñez, S. Recent developments on the catalytic technologies for the transformation of biomass into biofuels: A patent survey. Renew. Sustain. Energy Rev. 2015, 51, 273–287. [Google Scholar] [CrossRef]
  31. He, M.; Xiao, B.; Liu, S.; Guo, X.; Luo, S.; Xu, Z.; Feng, Y.; Hu, Z. Hydrogen-rich gas from catalytic steam gasification of municipal solid waste (MSW): Influence of steam to MSW ratios and weight hourly space velocity on gas production and composition. Int. J. Hydrog. Energy 2009, 34, 2174–2183. [Google Scholar] [CrossRef]
  32. Wang, J.; Cheng, G.; You, Y.; Xiao, B.; Liu, S.; He, P.; Guo, D.; Guo, X.; Zhang, G. Hydrogen-rich gas production by steam gasification of municipal solid waste (MSW) using NiO supported on modified dolomite. Int. J. Hydrog. Energy 2012, 37, 6503–6510. [Google Scholar] [CrossRef]
  33. Zhang, Q.; Dor, L.; Zhang, L.; Yang, W.; Blasiak, W. Performance analysis of municipal solid waste gasification with steam in a Plasma Gasification Melting reactor. Appl. Energy 2012, 98, 219–229. [Google Scholar] [CrossRef]
  34. Roche, E.; de Andrés, J.M.; Narros, A.; Rodríguez, M.E. Air and air-steam gasification of sewage sludge. The influence of dolomite and throughput in tar production and composition. Fuel 2014, 115, 54–61. [Google Scholar] [CrossRef]
  35. De Andrés, J.M.; Narros, A.; Rodríguez, M.E. Air-steam gasification of sewage sludge in a bubbling bed reactor: Effect of alumina as a primary catalyst. Fuel Process. Technol. 2011, 92, 433–440. [Google Scholar] [CrossRef] [Green Version]
  36. Gai, C.; Guo, Y.; Liu, T.; Peng, N.; Liu, Z. Hydrogen-Rich gas production by steam gasification of hydrochar derived from sewage sludge. Int. J. Hydrog. Energy 2016, 41, 3363–3372. [Google Scholar] [CrossRef]
  37. Xiao, X.; Le, D.D.; Li, L.; Meng, X.; Cao, J.; Morishita, K.; Takarada, T. Catalytic steam gasification of biomass in fluidized bed at low temperature: Conversion from livestock manure compost to hydrogen-rich syngas. Biomass Bioenergy 2010, 34, 1505–1512. [Google Scholar] [CrossRef]
  38. Yong, T.L.-K.; Matsumura, Y. Catalytic Gasification of Poultry Manure and Eucalyptus Wood Mixture in Supercritical Water. Ind. Eng. Chem. Res. 2012, 51, 5685–5690. [Google Scholar] [CrossRef]
  39. Ro, K.S.; Cantrell, K.; Elliott, D.; Hunt, P.G. Catalytic Wet Gasification of Municipal and Animal Wastes. Ind. Eng. Chem. Res. 2007, 46, 8839–8845. [Google Scholar] [CrossRef]
  40. Tushar, M.S.H.K.; Dutta, A.; Xu, C. Catalytic supercritical gasification of biocrude from hydrothermal liquefaction of cattle manure. Appl. Catal. B Environ. 2016, 189, 119–132. [Google Scholar] [CrossRef]
  41. Seif, S.; Tavakoli, O.; Fatemi, S.; Bahmanyar, H. Subcritical water gasification of beet-based distillery wastewater for hydrogen production. J. Supercrit. Fluids 2015, 104, 212–220. [Google Scholar] [CrossRef]
  42. Cao, C.; Guo, L.; Chen, Y.; Guo, S.; Lu, Y. Hydrogen production from supercritical water gasification of alkaline wheat straw pulping black liquor in continuous flow system. Int. J. Hydrog. Energy 2011, 36, 13528–13535. [Google Scholar] [CrossRef]
  43. Luo, S.; Xiao, B.; Hu, Z.; Liu, S.; Guo, X.; He, M. Hydrogen-rich gas from catalytic steam gasification of biomass in a fixed bed reactor: Influence of temperature and steam on gasification performance. Int. J. Hydrog. Energy 2009, 34, 2191–2194. [Google Scholar] [CrossRef]
  44. Wu, W.; Kawamoto, K.; Kuramochi, H. Hydrogen-rich synthesis gas production from waste wood via gasification and reforming technology for fuel cell application. J. Mater Cycles Waste Manag. 2006, 8, 70–77. [Google Scholar] [CrossRef]
  45. Ge, H.; Guo, W.; Shen, L.; Song, T.; Xiao, J. Experimental investigation on biomass gasification using chemical looping in a batch reactor and a continuous dual reactor. Chem. Eng. J. 2016, 286, 689–700. [Google Scholar] [CrossRef]
  46. Jin, H.; Lu, Y.; Guo, L.; Zhang, X.; Pei, A. Hydrogen Production by Supercritical Water Gasification of Biomass with Homogeneous and Heterogeneous Catalyst. Adv. Condens. Matter Phys. 2014, 2014, 160565. [Google Scholar] [CrossRef] [Green Version]
  47. Acelas, N.Y.; López, D.P.; Brilman, D.W.F.; Kersten, S.R.A.; Kootstra, A.M.J. Supercritical water gasification of sewage sludge: Gas production and phosphorus recovery. Bioresour. Technol. 2014, 174, 167–175. [Google Scholar] [CrossRef]
  48. Sawai, O.; Nunoura, T.; Yamamoto, K. Supercritical water gasification of sewage sludge using bench-scale batch reactor: Advantages and drawbacks. J. Mater Cycles Waste Manag. 2014, 16, 82–92. [Google Scholar] [CrossRef]
  49. Basu, P. (Ed.) Production of Synthetic Fuels and Chemicals from Biomass. In Biomass Gasification and Pyrolysis; Academic Press: Boston, MA, USA, 2010; Chapter 9; pp. 301–323. [Google Scholar]
  50. Higman, C.; van der Burgt, M. (Eds.) The Thermodynamics of Gasification. In Gasification; Gulf Professional Publishing: Burlington, VT, USA, 2003; Chapter 2; pp. 9–28. [Google Scholar]
  51. Maitlo, G.; Unar, I.N.; Mahar, R.B.; Brohi, K.M. Numerical simulation of lignocellulosic biomass gasification in concentric tube entrained flow gasifier through computational fluid dynamics. Energy Exploar. Exploit. 2019, 37, 1073–1097. [Google Scholar] [CrossRef] [Green Version]
  52. Maitlo, G.; Unar, I.N.; Shah, S.A.K. Thermogravimetric analysis of Pakistani biomasses using nitrogen and oxygen as a carrier gas. Chem. Pap. 2019, 73, 601–609. [Google Scholar] [CrossRef]
  53. Niu, M.; Huang, Y.; Jin, B.; Wang, X. Oxygen Gasification of Municipal Solid Waste in a Fixed-bed Gasifier. Chin. J. Chem. Eng. 2014, 22, 1021–1026. [Google Scholar] [CrossRef]
  54. Gil, J.; Corella, J.; Aznar, M.A.P.; Caballero, M.A. Biomass gasification in atmospheric and bubbling fluidized bed: Effect of the type of gasifying agent on the product distribution. Biomass Bioenergy 1999, 17, 389–403. [Google Scholar] [CrossRef]
  55. García, L.; Salvador, M.L.; Arauzo, J.; Bilbao, R. Catalytic Steam Gasification of Pine Sawdust. Effect of Catalyst Weight/Biomass Flow Rate and Steam/Biomass Ratios on Gas Production and Composition. Energy Fuels 1999, 13, 851–859. [Google Scholar] [CrossRef]
  56. Hernández, J.J.; Aranda, G.; Barba, J.; Mendoza, J.M. Effect of steam content in the air–steam flow on biomass entrained flow gasification. Fuel Process. Technol. 2012, 99, 43–55. [Google Scholar] [CrossRef]
  57. Rapagnà, S.; Jand, N.; Kiennemann, A.; Foscolo, P.U. Steam-Gasification of biomass in a fluidised-bed of olivine particles. Biomass Bioenergy 2000, 19, 187–197. [Google Scholar] [CrossRef]
  58. Li, X.; Li, C.-Z. Volatilisation and catalytic effects of alkali and alkaline earth metallic species during the pyrolysis and gasification of Victorian brown coal. Part VIII. Catalysis and changes in char structure during gasification in steam. Fuel 2006, 85, 1518–1525. [Google Scholar] [CrossRef]
  59. Lucas, C.; Szewczyk, D.; Blasiak, W.; Mochida, S. High-Temperature air and steam gasification of densified biofuels. Biomass Bioenergy 2004, 27, 563–575. [Google Scholar] [CrossRef]
  60. Navarro, R.M.; Peña, M.A.; Fierro, J.L.G. Hydrogen Production Reactions from Carbon Feedstocks:  Fossil Fuels and Biomass. Chem. Rev. 2007, 107, 3952–3991. [Google Scholar] [CrossRef] [PubMed]
  61. Hamelinck, C.N.; Faaij, A.P. Future prospects for production of methanol and hydrogen from biomass. J. Power Sources 2002, 111, 1–22. [Google Scholar] [CrossRef]
  62. Mohammed, M.A.A.; Salmiaton, A.; Wan Azlina, W.A.K.G.; Mohammad Amran, M.S.; Fakhru’l-Razi, A. Air gasification of empty fruit bunch for hydrogen-rich gas production in a fluidized-bed reactor. Energy Convers. Manag. 2011, 52, 1555–1561. [Google Scholar] [CrossRef]
  63. Xiao, R.; Jin, B.; Zhou, H.; Zhong, Z.; Zhang, M. Air gasification of polypropylene plastic waste in fluidized bed gasifier. Energy Convers. Manag. 2007, 48, 778–786. [Google Scholar] [CrossRef]
  64. Hamad, M.A.; Radwan, A.M.; Heggo, D.A.; Moustafa, T. Hydrogen rich gas production from catalytic gasification of biomass. Renew. Energy 2016, 85, 1290–1300. [Google Scholar] [CrossRef]
  65. Gao, N.; Li, A.; Quan, C.; Qu, Y.; Mao, L. Characteristics of hydrogen-rich gas production of biomass gasification with porous ceramic reforming. Int. J. Hydrog. Energy 2012, 37, 9610–9618. [Google Scholar] [CrossRef]
  66. Zhao, Y.; Sun, S.; Zhou, H.; Sun, R.; Tian, H.; Luan, J.; Qian, J. Experimental study on sawdust air gasification in an entrained-flow reactor. Fuel Process. Technol. 2010, 91, 910–914. [Google Scholar] [CrossRef]
  67. Manyà, J.J.; Sánchez, J.L.; Ábrego, J.; Gonzalo, A.; Arauzo, J. Influence of gas residence time and air ratio on the air gasification of dried sewage sludge in a bubbling fluidised bed. Fuel 2006, 85, 2027–2033. [Google Scholar] [CrossRef]
  68. Mansaray, K.G.; Ghaly, A.E.; Al-Taweel, A.M.; Hamdullahpur, F.; Ugursal, V.I. Air gasification of rice husk in a dual distributor type fluidized bed gasifier. Biomass Bioenergy 1999, 17, 315–332. [Google Scholar] [CrossRef]
  69. Islam, M.W. A review of dolomite catalyst for biomass gasification tar removal. Fuel 2020, 267, 117095. [Google Scholar] [CrossRef]
  70. Guan, Y.; Luo, S.; Liu, S.; Xiao, B.; Cai, L. Steam catalytic gasification of municipal solid waste for producing tar-free fuel gas. Int. J. Hydrog. Energy 2009, 34, 9341–9346. [Google Scholar] [CrossRef]
  71. Hu, M.; Guo, D.; Ma, C.; Hu, Z.; Zhang, B.; Xiao, B.; Luo, S.; Wang, J. Hydrogen-rich gas production by the gasification of wet MSW (municipal solid waste) coupled with carbon dioxide capture. Energy 2015, 90, 857–863. [Google Scholar] [CrossRef]
  72. Dong, L.; Wu, C.; Ling, H.; Shi, J.; Williams, P.T.; Huang, J. Promoting hydrogen production and minimizing catalyst deactivation from the pyrolysis-catalytic steam reforming of biomass on nanosized NiZnAlOx catalysts. Fuel 2017, 188, 610–620. [Google Scholar] [CrossRef]
  73. Ma, Z.; Zhang, S.-P.; Xie, D.-Y.; Yan, Y.-J. A novel integrated process for hydrogen production from biomass. Int. J. Hydrog. Energy 2014, 39, 1274–1279. [Google Scholar] [CrossRef]
  74. Yao, D.; Wu, C.; Yang, H.; Hu, Q.; Nahil, M.A.; Chen, H.; Williams, P.T. Hydrogen production from catalytic reforming of the aqueous fraction of pyrolysis bio-oil with modified Ni–Al catalysts. Int. J. Hydrog. Energy 2014, 39, 14642–14652. [Google Scholar] [CrossRef] [Green Version]
  75. Cao, J.-P.; Liu, T.-L.; Ren, J.; Zhao, X.-Y.; Wu, Y.; Wang, J.-X.; Ren, X.-Y.; Wei, X.-Y. Preparation and characterization of nickel loaded on resin char as tar reforming catalyst for biomass gasification. J. Anal. Appl. Pyrolysis 2017, 127, 82–90. [Google Scholar] [CrossRef]
  76. Ren, J.; Cao, J.-P.; Zhao, X.-Y.; Wei, F.; Liu, T.-L.; Fan, X.; Zhao, Y.-P.; Wei, X.-Y. Preparation of high-dispersion Ni/C catalyst using modified lignite as carbon precursor for catalytic reforming of biomass volatiles. Fuel 2017, 202, 345–351. [Google Scholar] [CrossRef]
  77. Zhao, X.-Y.; Ren, J.; Cao, J.-P.; Wei, F.; Zhu, C.; Fan, X.; Zhao, Y.-P.; Wei, X.-Y. Catalytic Reforming of Volatiles from Biomass Pyrolysis for Hydrogen-Rich Gas Production over Limonite Ore. Energy Fuels 2017, 31, 4054–4060. [Google Scholar] [CrossRef]
  78. Fu, P.; Yi, W.; Li, Z.; Bai, X.; Zhang, A.; Li, Y.; Li, Z. Investigation on hydrogen production by catalytic steam reforming of maize stalk fast pyrolysis bio-oil. Int. J. Hydrog. Energy 2014, 39, 13962–13971. [Google Scholar] [CrossRef]
  79. Remón, J.; Broust, F.; Valette, J.; Chhiti, Y.; Alava, I.; Fernandez-Akarregi, A.R.; Arauzo, J.; Garcia, L. Production of a hydrogen-rich gas from fast pyrolysis bio-oils: Comparison between homogeneous and catalytic steam reforming routes. Int. J. Hydrog. Energy 2014, 39, 171–182. [Google Scholar] [CrossRef]
  80. Remiro, A.; Valle, B.; Aguayo, A.T.; Bilbao, J.; Gayubo, A.G. Operating conditions for attenuating Ni/La2O3–αAl2O3 catalyst deactivation in the steam reforming of bio-oil aqueous fraction. Fuel Process. Technol. 2013, 115, 222–232. [Google Scholar] [CrossRef]
  81. Bimbela, F.; Oliva, M.; Ruiz, J.; García, L.; Arauzo, J. Hydrogen production via catalytic steam reforming of the aqueous fraction of bio-oil using nickel-based coprecipitated catalysts. Int. J. Hydrog. Energy 2013, 38, 14476–14487. [Google Scholar] [CrossRef]
  82. Li, W.; Ren, J.; Zhao, X.-Y.; Takarada, T. H and Syngas Production From Catalytic Cracking of Pig Manure and Compost Pyrolysis Vapor Over Ni-Based Catalysts. Pol. J. Chem. Technol. 2018, 20, 8–14. [Google Scholar] [CrossRef] [Green Version]
  83. Li, H.; Xu, Q.; Xue, H.; Yan, Y. Catalytic reforming of the aqueous phase derived from fast-pyrolysis of biomass. Renew. Energy 2009, 34, 2872–2877. [Google Scholar] [CrossRef]
  84. Yan, C.-F.; Cheng, F.-F.; Hu, R.-R. Hydrogen production from catalytic steam reforming of bio-oil aqueous fraction over Ni/CeO2–ZrO2 catalysts. Int. J. Hydrog. Energy 2010, 35, 11693–11699. [Google Scholar] [CrossRef]
  85. Wu, C.; Wang, L.; Williams, P.T.; Shi, J.; Huang, J. Hydrogen production from biomass gasification with Ni/MCM-41 catalysts: Influence of Ni content. Appl. Catal. B Environ. 2011, 108, 6–13. [Google Scholar] [CrossRef]
  86. Wu, C.; Huang, Q.; Sui, M.; Yan, Y.; Wang, F. Hydrogen production via catalytic steam reforming of fast pyrolysis bio-oil in a two-stage fixed bed reactor system. Fuel Process. Technol. 2008, 89, 1306–1316. [Google Scholar] [CrossRef]
  87. Warnecke, R. Gasification of biomass: Comparison of fixed bed and fluidized bed gasifier. Biomass Bioenergy 2000, 18, 489–497. [Google Scholar] [CrossRef]
  88. Mandal, S.; Daggupati, S.; Majhi, S.; Thakur, S.; Bandyopadhyay, R.; Das, A.K. Catalytic Gasification of Biomass in Dual-Bed Gasifier for Producing Tar-Free Syngas. Energy Fuels 2019, 33, 2453–2466. [Google Scholar] [CrossRef]
  89. Chen, W.-H.; Chen, C.-J.; Hung, C.-I.; Shen, C.-H.; Hsu, H.-W. A comparison of gasification phenomena among raw biomass, torrefied biomass and coal in an entrained-flow reactor. Appl. Energy 2013, 112, 421–430. [Google Scholar] [CrossRef]
  90. Janajreh, I.; Raza, S.S.; Valmundsson, A.S. Plasma gasification process: Modeling, simulation and comparison with conventional air gasification. Energy Convers. Manag. 2013, 65, 801–809. [Google Scholar] [CrossRef]
  91. Salam, M.A.; Ahmed, K.; Akter, N.; Hossain, T.; Abdullah, B. A review of hydrogen production via biomass gasification and its prospect in Bangladesh. Int. J. Hydrog. Energy 2018, 43, 14944–14973. [Google Scholar] [CrossRef]
  92. Ren, J.; Cao, J.-P.; Zhao, X.-Y.; Yang, F.-L.; Wei, X.-Y. Recent advances in syngas production from biomass catalytic gasification: A critical review on reactors, catalysts, catalytic mechanisms and mathematical models. Renew. Sustain. Energy Rev. 2019, 116, 109426. [Google Scholar] [CrossRef]
  93. Buragohain, B.; Mahanta, P.; Moholkar, V.S. Biomass gasification for decentralized power generation: The Indian perspective. Renew. Sustain. Energy Rev. 2010, 14, 73–92. [Google Scholar] [CrossRef]
  94. Ismail, T.M.; El-Salam, M.A. Parametric studies on biomass gasification process on updraft gasifier high temperature air gasification. Appl. Therm. Eng. 2017, 112, 1460–1473. [Google Scholar] [CrossRef]
  95. Kihedu, J.H.; Yoshiie, R.; Naruse, I. Performance indicators for air and air–steam auto-thermal updraft gasification of biomass in packed bed reactor. Fuel Process. Technol. 2016, 141, 93–98. [Google Scholar] [CrossRef]
  96. Susastriawan, A.A.P.; Saptoadi, H. Purnomo, Small-scale downdraft gasifiers for biomass gasification: A review. Renew. Sustain. Energy Rev. 2017, 76, 989–1003. [Google Scholar] [CrossRef]
  97. Kuo, P.-C.; Wu, W.; Chen, W.-H. Gasification performances of raw and torrefied biomass in a downdraft fixed bed gasifier using thermodynamic analysis. Fuel 2014, 117, 1231–1241. [Google Scholar] [CrossRef]
  98. Nwakaire, J.; Ugwuishiwu, B. Development of a Natural Cross Draft Gasifier Stove for Application in Rural Communities in Sub-Saharan Africa. J. Appl. Sci. 2015, 15, 1149–1157. [Google Scholar] [CrossRef] [Green Version]
  99. Udomsirichakorn, J.; Basu, P.; Salam, P.A.; Acharya, B. Effect of CaO on tar reforming to hydrogen-enriched gas with in-process CO2 capture in a bubbling fluidized bed biomass steam gasifier. Int. J. Hydrog. Energy 2013, 38, 14495–14504. [Google Scholar] [CrossRef]
  100. Kirnbauer, F.; Wilk, V.; Kitzler, H.; Kern, S.; Hofbauer, H. The positive effects of bed material coating on tar reduction in a dual fluidized bed gasifier. Fuel 2012, 95, 553–562. [Google Scholar] [CrossRef]
  101. Corella, J.; Toledo, J.M.; Molina, G. A Review on Dual Fluidized-Bed Biomass Gasifiers. Ind. Eng. Chem. Res. 2007, 46, 6831–6839. [Google Scholar] [CrossRef]
  102. Couhert, C.; Salvador, S.; Commandré, J.M. Impact of torrefaction on syngas production from wood. Fuel 2009, 88, 2286–2290. [Google Scholar] [CrossRef] [Green Version]
  103. Chhiti, Y.; Peyrot, M.; Salvador, S. Soot formation and oxidation during bio-oil gasification: Experiments and modeling. J. Energy Chem. 2013, 22, 701–709. [Google Scholar] [CrossRef] [Green Version]
  104. Zhang, Q.; Dor, L.; Fenigshtein, D.; Yang, W.; Blasiak, W. Gasification of municipal solid waste in the Plasma Gasification Melting process. Appl. Energy 2012, 90, 106–112. [Google Scholar] [CrossRef]
  105. Seo, Y.-C.; Alam, M.T.; Yang, W.-S. Gasification of municipal solid waste. In Gasification Low-Grade Feedstock; InTech Open: London, UK, 2018. [Google Scholar]
  106. Mountouris, A.; Voutsas, E.; Tassios, D. Solid waste plasma gasification: Equilibrium model development and exergy analysis. Energy Convers. Manag. 2006, 47, 1723–1737. [Google Scholar] [CrossRef]
  107. Chen, G.; Yao, J.; Yang, H.; Yan, B.; Chen, H. Steam gasification of acid-hydrolysis biomass CAHR for clean syngas production. Bioresour. Technol. 2015, 179, 323–330. [Google Scholar] [CrossRef] [PubMed]
  108. Udomsirichakorn, J.; Salam, P.A. Review of hydrogen-enriched gas production from steam gasification of biomass: The prospect of CaO-based chemical looping gasification. Renew. Sustain. Energy Rev. 2014, 30, 565–579. [Google Scholar] [CrossRef]
  109. Nipattummakul, N.; Ahmed, I.I.; Kerdsuwan, S.; Gupta, A.K. Hydrogen and syngas production from sewage sludge via steam gasification. Int. J. Hydrog. Energy 2010, 35, 11738–11745. [Google Scholar] [CrossRef]
  110. Hu, M.; Gao, L.; Chen, Z.; Ma, C.; Zhou, Y.; Chen, J.; Ma, S.; Laghari, M.; Xiao, B.; Zhang, B.; et al. Syngas production by catalytic in-situ steam co-gasification of wet sewage sludge and pine sawdust. Energy Convers. Manag. 2016, 111, 409–416. [Google Scholar] [CrossRef]
  111. Wei, L.; Xu, S.; Zhang, L.; Liu, C.; Zhu, H.; Liu, S. Steam gasification of biomass for hydrogen-rich gas in a free-fall reactor. Int. J. Hydrog. Energy 2007, 32, 24–31. [Google Scholar] [CrossRef]
  112. Wu, C.; Wang, Z.; Dupont, V.; Huang, J.; Williams, P.T. Nickel-catalysed pyrolysis/gasification of biomass components. J. Anal. Appl. Pyrolysis 2013, 99, 143–148. [Google Scholar] [CrossRef]
  113. Veraa, M.J.; Bell, A.T. Effect of alkali metal catalysts on gasification of coal char. Fuel 1978, 57, 194–200. [Google Scholar] [CrossRef]
  114. Li, J.; Liao, S.; Dan, W.; Jia, K.; Zhou, X. Experimental study on catalytic steam gasification of municipal solid waste for bioenergy production in a combined fixed bed reactor. Biomass Bioenergy 2012, 46, 174–180. [Google Scholar] [CrossRef]
  115. Abuadala, A.; Dincer, I. Investigation of a multi-generation system using a hybrid steam biomass gasification for hydrogen, power and heat. Int. J. Hydrog. Energy 2010, 35, 13146–13157. [Google Scholar] [CrossRef]
  116. Tursun, Y.; Xu, S.; Wang, C.; Xiao, Y.; Wang, G. Steam co-gasification of biomass and coal in decoupled reactors. Fuel Process. Technol. 2016, 141, 61–67. [Google Scholar] [CrossRef]
  117. Kuo, P.-C.; Wu, W. Design, Optimization and Energetic Efficiency of Producing Hydrogen-Rich Gas from Biomass Steam Gasification. Energies 2015, 8, 94–110. [Google Scholar] [CrossRef]
  118. Pfeifer, C.; Hofbauer, H. Development of catalytic tar decomposition downstream from a dual fluidized bed biomass steam gasifier. Powder Technol. 2008, 180, 9–16. [Google Scholar] [CrossRef]
  119. Yan, F.; Luo, S.-Y.; Hu, Z.-Q.; Xiao, B.; Cheng, G. Hydrogen-rich gas production by steam gasification of char from biomass fast pyrolysis in a fixed-bed reactor: Influence of temperature and steam on hydrogen yield and syngas composition. Bioresour. Technol. 2010, 101, 5633–5637. [Google Scholar] [CrossRef]
  120. Milne, T.A.; Evans, R.J.; Abatzaglou, N. Biomass Gasifier “Tars”: Their Nature, Formation, and Conversion; National Renewable Energy Laboratory: Golden, CO, USA, 1998. [Google Scholar]
  121. Fuentes-Cano, D.; Gómez-Barea, A.; Nilsson, S.; Ollero, P. The influence of temperature and steam on the yields of tar and light hydrocarbon compounds during devolatilization of dried sewage sludge in a fluidized bed. Fuel 2013, 108, 341–350. [Google Scholar] [CrossRef]
  122. Zhang, Y.; Kajitani, S.; Ashizawa, M.; Oki, Y. Tar destruction and coke formation during rapid pyrolysis and gasification of biomass in a drop-tube furnace. Fuel 2010, 89, 302–309. [Google Scholar] [CrossRef]
  123. Jess, A. Mechanisms and kinetics of thermal reactions of aromatic hydrocarbons from pyrolysis of solid fuels. Fuel 1996, 75, 1441–1448. [Google Scholar] [CrossRef]
  124. Watson, J.; Zhang, Y.; Si, B.; Chen, W.-T.; de Souza, R. Gasification of biowaste: A critical review and outlooks. Renew. Sustain. Energy Rev. 2018, 83, 1–17. [Google Scholar] [CrossRef]
  125. Hosseini, M.; Dincer, I.; Rosen, M.A. Steam and air fed biomass gasification: Comparisons based on energy and exergy. Int. J. Hydrog. Energy 2012, 37, 16446–16452. [Google Scholar] [CrossRef]
  126. Wu, Y.; Yang, W.; Blasiak, W. Energy and Exergy Analysis of High Temperature Agent Gasification of Biomass. Energies 2014, 7, 2107–2122. [Google Scholar] [CrossRef]
  127. Song, G.; Chen, L.; Xiao, J.; Shen, L. Exergy evaluation of biomass steam gasification via interconnected fluidized beds. Int. J. Energy Res. 2013, 37, 1743–1751. [Google Scholar] [CrossRef]
  128. Heidenreich, S.; Foscolo, P.U. New concepts in biomass gasification. Prog. Energy Combust. Sci. 2015, 46, 72–95. [Google Scholar] [CrossRef]
  129. Dudyński, M.; van Dyk, J.C.; Kwiatkowski, K.; Sosnowska, M. Biomass gasification: Influence of torrefaction on syngas production and tar formation. Fuel Process. Technol. 2015, 131, 203–212. [Google Scholar] [CrossRef]
  130. Valderrama Rios, M.L.; González, A.M.; Lora, E.E.S.; del Olmo, O.A.A. Reduction of tar generated during biomass gasification: A review. Biomass Bioenergy 2018, 108, 345–370. [Google Scholar] [CrossRef]
  131. Ren, J.; Liu, Y.-L.; Zhao, X.-Y.; Cao, J.-P. Biomass thermochemical conversion: A review on tar elimination from biomass catalytic gasification. J. Energy Inst. 2020, 93, 1083–1098. [Google Scholar] [CrossRef]
  132. De Caprariis, B.; Bassano, C.; Deiana, P.; Palma, V.; Petrullo, A.; Scarsella, M.; De Filippis, P. Carbon dioxide reforming of tar during biomass gasification. Chem. Eng. Trans. 2014, 37, 97–102. [Google Scholar]
  133. Dahlquist, E. Technologies for Converting Biomass to Useful Energy: Combustion, Gasification, Pyrolysis, Torrefaction and Fermentation; CRC Press: Boca Raton, FL, USA, 2013. [Google Scholar]
  134. Pio, D.T.; Tarelho, L.A.C.; Pinto, R.G.; Matos, M.A.A.; Frade, J.R.; Yaremchenko, A.; Mishra, G.S.; Pinto, P.C.R. Low-cost catalysts for in-situ improvement of producer gas quality during direct gasification of biomass. Energy 2018, 165, 442–454. [Google Scholar] [CrossRef]
  135. Devi, L.; Ptasinski, K.J.; Janssen, F.J.J.G. A review of the primary measures for tar elimination in biomass gasification processes. Biomass Bioenergy 2003, 24, 125–140. [Google Scholar] [CrossRef]
  136. Boerrigter, H.; Van Paasen, S.; Bergman, P.; Könemann, J.; Emmen, R.; Wijnands, A. OLGA tar Removal Technology; ECN-C--05–009; Energy Research Centre of The Netherlands: Petten, The Netherlands, 2005. [Google Scholar]
  137. Basu, P. Biomass Gasification and Pyrolysis: Practical Design and Theory; Academic Press: Cambridge, MA, USA, 2010. [Google Scholar]
  138. Göransson, K.; Söderlind, U.; He, J.; Zhang, W. Review of syngas production via biomass DFBGs. Renew. Sustain. Energy Rev. 2011, 15, 482–492. [Google Scholar] [CrossRef]
  139. Pfeifer, C.; Koppatz, S.; Hofbauer, H. Steam gasification of various feedstocks at a dual fluidised bed gasifier: Impacts of operation conditions and bed materials. Biomass Convers. Biorefin. 2011, 1, 39–53. [Google Scholar] [CrossRef]
  140. Mednikov, A.S. A Review of Technologies for Multistage Wood Biomass Gasification. Therm. Eng. 2018, 65, 531–546. [Google Scholar] [CrossRef]
  141. Fabry, F.; Rehmet, C.; Rohani, V.; Fulcheri, L. Waste Gasification by Thermal Plasma: A Review. Waste Biomass Valoriz. 2013, 4, 421–439. [Google Scholar] [CrossRef] [Green Version]
  142. Asadullah, M. Barriers of commercial power generation using biomass gasification gas: A review. Renew. Sustain. Energy Rev. 2014, 29, 201–215. [Google Scholar] [CrossRef]
  143. Asadullah, M. Biomass gasification gas cleaning for downstream applications: A comparative critical review. Renew. Sustain. Energy Rev. 2014, 40, 118–132. [Google Scholar] [CrossRef]
  144. Raman, P.; Ram, N.K.; Gupta, R. A dual fired downdraft gasifier system to produce cleaner gas for power generation: Design, development and performance analysis. Energy 2013, 54, 302–314. [Google Scholar] [CrossRef]
  145. Couto, N.; Rouboa, A.; Silva, V.; Monteiro, E.; Bouziane, K. Influence of the Biomass Gasification Processes on the Final Composition of Syngas. Energy Procedia 2013, 36, 596–606. [Google Scholar] [CrossRef] [Green Version]
  146. Siedlecki, M.; De Jong, W.; Verkooijen, A.H.M. Fluidized Bed Gasification as a Mature And Reliable Technology for the Production of Bio-Syngas and Applied in the Production of Liquid Transportation Fuels—A Review. Energies 2011, 4, 389–434. [Google Scholar] [CrossRef] [Green Version]
  147. Mondal, P.; Dang, G.S.; Garg, M.O. Syngas production through gasification and cleanup for downstream applications—Recent developments. Fuel Process. Technol. 2011, 92, 1395–1410. [Google Scholar] [CrossRef]
  148. Centeno, F.; Mahkamov, K.; Silva Lora, E.E.; Andrade, R.V. Theoretical and experimental investigations of a downdraft biomass gasifier-spark ignition engine power system. Renew. Energy 2012, 37, 97–108. [Google Scholar] [CrossRef]
  149. Siedlecki, M.; De Jong, W. Biomass gasification as the first hot step in clean syngas production process—Gas quality optimization and primary tar reduction measures in a 100 kW thermal input steam–oxygen blown CFB gasifier. Biomass Bioenergy 2011, 35, S40–S62. [Google Scholar] [CrossRef]
  150. Qin, K.; Jensen, P.A.; Lin, W.; Jensen, A.D. Biomass Gasification Behavior in an Entrained Flow Reactor: Gas Product Distribution and Soot Formation. Energy Fuels 2012, 26, 5992–6002. [Google Scholar] [CrossRef] [Green Version]
  151. Donaj, P.; Amovic, M.; Moner, B.; Engvall, K. Flexibility and robustness of WoodRoll system—Tests results from a 500 kW plant. In Proceedings of the 1st International Conference on Renewable EnergyGas Technology (REGATEC 2014), Malmö, Sweden, 22–23 May 2014. [Google Scholar]
  152. Barisano, D. Biomass Gasification for Power Production: Country Report-Italy; Italian National Agency for New Technologies, Energy and Sustainable Economic Development: Rome, Italy, 2017. [Google Scholar]
  153. Brynda, J.; Skoblia, S.; Beňo, Z.; Pohořelý, M.; Moško, J. Application of staged biomass gasification for combined heat and power production. In Proceedings of the 1st International Conference on Environmental Technology and Innovations, Ho Chi Minh City, Vietnam, 23–25 November 2016; CRC Press: Boca Raton, FL, USA, 2016; p. 143. [Google Scholar]
  154. Skoblia, S.; Beňo, Z.; Brynda, J.; Moško, J.; Pohořely, M. Experience with Operation of Multi-Stage (Two-Stage) Fixed-Bed Gasifiers in the Czech and Slovak Republic. In Proceedings of the 8th International Fieburg Conference on IGCC and XtL Technologies, Köln, Germany, 12–16 June 2016. [Google Scholar]
  155. Henriksen, U.; Ahrenfeldt, J.; Jensen, T.K.; Gøbel, B.; Bentzen, J.D.; Hindsgaul, C.; Sørensen, L.H. The design, construction and operation of a 75 kW two-stage gasifier. Energy 2006, 31, 1542–1553. [Google Scholar] [CrossRef] [Green Version]
  156. Bentzen, J.D.; Hummelshøj, R.; Henriksen, U.; Gøbel, B.; Ahrenfelt, J.; Elmegaard, B. Upscale of the two-stage gasification process. In Proceedings of the 2nd World Conference on Biomass, Rome, Italy, 10–14 May 2004; pp. 1004–1007. [Google Scholar]
  157. Bolhar-Nordenkampf, M.; Rauch, R.; Bosch, K.; Aichernig, C.; Hofbauer, H. Biomass CHP plant Güssing—Using gasification for power generation. In Proceedings of the 2nd Regional Conference on Energy Technology towards a Clean Environment, Phuket, Thailand, 12–14 February 2003. [Google Scholar]
  158. Gadsbøll, R.Ø.; Sárossy, Z.; Jørgensen, L.; Ahrenfeldt, J.; Henriksen, U.B. Oxygen-blown operation of the TwoStage Viking gasifier. Energy 2018, 158, 495–503. [Google Scholar] [CrossRef]
  159. Teh, J.S.; Teoh, Y.H.; How, H.G.; Le, T.D.; Jason, Y.J.J.; Nguyen, H.T.; Loo, D.L. The Potential of Sustainable Biomass Producer Gas as a Waste-to-Energy Alternative in Malaysia. Sustainability 2021, 13, 3877. [Google Scholar] [CrossRef]
  160. Hoque, M.; Rashid, F.; Aziz, M. Gasification and power generation characteristics of rice husk, sawdust, and coconut shell using a fixed-bed downdraft gasifier. Sustainability 2021, 13, 2027. [Google Scholar] [CrossRef]
  161. Konda, R.E.; Sulaiman, S.A.; Ariwahjoedi, B. Design and development of laboratory scale updraft gasifier for gasification of oil palm fronds. Asian J. Sci. Res. 2014, 7, 341. [Google Scholar] [CrossRef] [Green Version]
  162. Ahmed, I.I.; Gupta, A.K. Pyrolysis and gasification of food waste: Syngas characteristics and char gasification kinetics. Appl. Energy 2010, 87, 101–108. [Google Scholar] [CrossRef]
  163. Seo, D.K.; Lee, S.K.; Kang, M.W.; Hwang, J.; Yu, T.-U. Gasification reactivity of biomass chars with CO2. Biomass and Bioenergy 2010, 34, 1946–1953. [Google Scholar] [CrossRef]
  164. Umeki, K.; Roh, S.A.; Min, T.J.; Namioka, T.; Yoshikawa, K. A simple expression for the apparent reaction rate of large wood char gasification with steam. Bioresour. Technol. 2010, 101, 4187–4192. [Google Scholar] [CrossRef]
  165. Matsumoto, K.; Takeno, K.; Ichinose, T.; Ogi, T.; Nakanishi, M. Gasification reaction kinetics on biomass char obtained as a by-product of gasification in an entrained-flow gasifier with steam and oxygen at 900–1000°C. Fuel 2009, 88, 519–527. [Google Scholar] [CrossRef]
  166. Okumura, Y.; Hanaoka, T.; Sakanishi, K. Effect of pyrolysis conditions on gasification reactivity of woody biomass-derived char. Proceedings of the Combustion Institute 2009, 32, 2013–2020. [Google Scholar] [CrossRef]
  167. Jiménez, S.; Remacha, P.; Ballesteros, J.C.; Giménez, A.; Ballester, J. Kinetics of devolatilization and oxidation of a pulverized biomass in an entrained flow reactor under realistic combustion conditions. Combustion and Flame 2008, 152, 588–603. [Google Scholar] [CrossRef] [Green Version]
  168. Fermoso, J.; Arias, B.; Pevida, C.; Plaza, M.; Rubiera, F.; Pis, J. Kinetic models comparison for steam gasification of different nature fuel chars. J. Therm. Anal. Calorim. 2008, 91, 779–786. [Google Scholar] [CrossRef] [Green Version]
  169. Guizani, C.; Escudero Sanz, F.J.; Salvador, S. The gasification reactivity of high-heating-rate chars in single and mixed atmospheres of H2O and CO2. Fuel 2013, 108, 812–823. [Google Scholar] [CrossRef] [Green Version]
  170. Lin, L.; Strand, M. Investigation of the intrinsic CO2 gasification kinetics of biomass char at medium to high temperatures. Appl. Energy 2013, 109, 220–228. [Google Scholar] [CrossRef]
  171. Kirtania, K.; Joshua, J.; Kassim, M.A.; Bhattacharya, S. Comparison of CO2 and steam gasification reactivity of algal and woody biomass chars. Fuel Processing Technol. 2014, 117, 44–52. [Google Scholar] [CrossRef]
  172. Bryan Woodruff, R.; Weimer, A.W. A novel technique for measuring the kinetics of high-temperature gasification of biomass char with steam. Fuel 2013, 103, 749–757. [Google Scholar] [CrossRef]
  173. Hanaoka, T.; Sakanishi, K.; Okumura, Y. The effect of N2/CO2/O2 content and pressure on characteristics and CO2 gasification behavior of biomass-derived char. Fuel Processing Technology 2012, 104, 287–294. [Google Scholar] [CrossRef]
  174. Irfan, M.F.; Arami-Niya, A.; Chakrabarti, M.H.; Wan Daud, W.M.A.; Usman, M.R. Kinetics of gasification of coal, biomass and their blends in air (N2/O2) and different oxy-fuel (O2/CO2) atmospheres. Energy 2012, 37, 665–672. [Google Scholar] [CrossRef]
  175. Mohammed, M.A.; Salmiaton, A.; Wan Azlina, W.A.; Mohamad Amran, M.S. Gasification of oil palm empty fruit bunches: A characterization and kinetic study. Bioresource technology 2012, 110, 628–636. [Google Scholar] [CrossRef]
  176. Konttinen, J.T.; Moilanen, A.; DeMartini, N.; Hupa, M. Carbon conversion predictor for fluidized bed gasification of biomass fuels—from TGA measurements to char gasification particle model. Biomass Conversion and Biorefinery 2012, 2, 265–274. [Google Scholar] [CrossRef]
  177. Nowicki, L.; Antecka, A.; Bedyk, T.; Stolarek, P.; Ledakowicz, S. The kinetics of gasification of char derived from sewage sludge. Journal of Thermal Analysis and Calorimetry 2011, 104, 693–700. [Google Scholar] [CrossRef] [Green Version]
  178. Ahmed, I.I.; Gupta, A.K. Kinetics of woodchips char gasification with steam and carbon dioxide. Applied Energy 2011, 88, 1613–1619. [Google Scholar] [CrossRef]
  179. Sakaguchi, M.; Watkinson, A.P.; Ellis, N. Steam gasification reactivity of char from rapid pyrolysis of bio-oil/char slurry. Fuel 2010, 89, 3078–3084. [Google Scholar] [CrossRef]
  180. Piatkowski, N.; Steinfeld, A. Reaction kinetics of the combined pyrolysis and steam-gasification of carbonaceous waste materials. Fuel 2010, 89, 1133–1140. [Google Scholar] [CrossRef]
Figure 1. Biomass conversion routes for achieving various products.
Figure 1. Biomass conversion routes for achieving various products.
Sustainability 14 02596 g001
Figure 2. Structure and characteristics of biomass concerning lignin, cellulose, and hemicellulose. Reprinted with permission from ref. [23]. Copyright 2018 Elsevier.
Figure 2. Structure and characteristics of biomass concerning lignin, cellulose, and hemicellulose. Reprinted with permission from ref. [23]. Copyright 2018 Elsevier.
Sustainability 14 02596 g002
Figure 3. (ah). Schematic presentation of different types of gasifiers. (a) Updraft [91]; (b) downdraft [91]; (c) crossdraft [92]; (d) bubbling [92]; (e) circulating [92]; (f) entrained flow [91]; (g) plasma [92]; (h) rotary kiln [92]. Reprinted Figure 3a,b,f with permission from ref. [91]. Copyright 2018 Elsevier. Reprinted Figure 3c,d,e,g,h with permission from ref. [92]. Copyright 2019 Elsevier.
Figure 3. (ah). Schematic presentation of different types of gasifiers. (a) Updraft [91]; (b) downdraft [91]; (c) crossdraft [92]; (d) bubbling [92]; (e) circulating [92]; (f) entrained flow [91]; (g) plasma [92]; (h) rotary kiln [92]. Reprinted Figure 3a,b,f with permission from ref. [91]. Copyright 2018 Elsevier. Reprinted Figure 3c,d,e,g,h with permission from ref. [92]. Copyright 2019 Elsevier.
Sustainability 14 02596 g003aSustainability 14 02596 g003b
Figure 4. Effect of S/B ratio on syngas composition during the gasification of biomass. Reprinted with permission from ref. [124]. Copyright 2018 Elsevier.
Figure 4. Effect of S/B ratio on syngas composition during the gasification of biomass. Reprinted with permission from ref. [124]. Copyright 2018 Elsevier.
Sustainability 14 02596 g004
Figure 5. Effect of S/B ratio of syngas production, LHV, CCE, and tar production for biowaste gasification. Reprinted with permission from ref. [124]. Copyright 2018 Elsevier.
Figure 5. Effect of S/B ratio of syngas production, LHV, CCE, and tar production for biowaste gasification. Reprinted with permission from ref. [124]. Copyright 2018 Elsevier.
Sustainability 14 02596 g005
Figure 6. S/B ratio and its effect on exergy efficiency during biowaste gasification. Reprinted with permission from ref. [124]. Copyright 2018 Elsevier.
Figure 6. S/B ratio and its effect on exergy efficiency during biowaste gasification. Reprinted with permission from ref. [124]. Copyright 2018 Elsevier.
Sustainability 14 02596 g006
Figure 7. CGE, CCE, and LHV of (a) raw bamboo, (b) torrefied bamboo at 250 °C, and (c) torrefied bamboo at 300 °C. Reprinted with permission from ref. [97]. Copyright 2014 Elsevier.
Figure 7. CGE, CCE, and LHV of (a) raw bamboo, (b) torrefied bamboo at 250 °C, and (c) torrefied bamboo at 300 °C. Reprinted with permission from ref. [97]. Copyright 2014 Elsevier.
Sustainability 14 02596 g007aSustainability 14 02596 g007b
Table 1. Different fuels, heating values and comparison of prices. Reprinted with permission from ref. [23]. Copyright 2018 Elsevier.
Table 1. Different fuels, heating values and comparison of prices. Reprinted with permission from ref. [23]. Copyright 2018 Elsevier.
FeedstockLower Heating ValueLHV (MJ/Kg)Estimated Price (USD/MJ)PricePrice USD/kg
Naphtha44.9 MJ/kg44.90.0120.52 USD/kg0.52
Coal (wet basis)22.7 MJ/kg22.70.0030.06 USD/kg0.06
Diesel35.9 MJ/L42.9 a0.0311.34 USD/L1.58
Syngas from biomass as FT uses10.0 MJ/m310.5 b0.0100.10 USD/m30.11
a Diesel density = 0.84 kg/L, b Syngas density = 0.95 kg/m3.
Table 2. List of main biomass gasification reactions. Reprinted with permission from [24]. Copyright 2013 Elsevier.
Table 2. List of main biomass gasification reactions. Reprinted with permission from [24]. Copyright 2013 Elsevier.
Main Reactions:
CH x O y + aH 2 +   wH 2 O + 3.76 aN 2 =   N 1 H 2 +   n 2 CO +   N 1 H 2 +   n 3 C 0 2 +   n 4 H 2 O +   n 5 CH 4 +   n 6 N 2 +   n 7 C   Char + Tar   (1)
Oxidation
1 C + 1 2 O 2 CO   (2)−111 MJ/kmolCarbon partial oxidation
2 CO + 1 2 O 2   CO 2   (3)−283 MJ/kmolCarbon monoxide oxidation
3 C +   O 2     CO 2 (4)−394 MJ/kmolCarbon oxidation
4 H 2 + 1 2 O 2   H 2 O (5)−242 MJ/kmolH2 oxidation
5 CH 4 + 1 / 2 O 2   CO + 2 H 2 −36 MJ/kmolCH4 partial oxidation
6 CH 4 + 2 O 2   CO 2 + 2 H 2 O (6)803 MJ/kmolOxidation
Gasification reactions involving steam:
7 CH 4 +   H 2 O     CO + 3 H 2   (7)206 MJ/kmolSteam methane reforming
8 CO +   H 2 O     CO 2 + H 2 −41 MJ/kmolWater–gas shift reaction
9 C +   H 2 O     CO +   H 2 131 MJ/kmolWater–gas reaction
10 C n H m + nH 2 O   nCO + n + m 2   H 2   (8)EndothermicSteam reforming
Gasification reactions involving H2:
11 C + 2 H 2     CH 4 (9)−75 MJ/kmolHydrogasification
12 CO + 3 H 2     CH 4 +   H 2 O (10)−227 MJ/kmolMethanation
13 2 CO + 2 H 2     CH 4 +   CO 2 (11)247 MJ/kmolMethanation
14   CO 2 + 4 H 2     CH 4 + 2 H 2 O 165 MJ/kmolMethanation
Gasification reactions involving carbon dioxide:
15 C +   CO 2     2 CO (12)172 MJ/kmolBoudouard reaction
16 C n H m + nCO 2   nCO + m 2 H 2 (14)EndothermicDry reforming
Decomposition reactions of tars and hydrocarbons:
17 pC x H y     qC n H m +   rH 2   (15)EndothermicDehydrogenation
18 C n H m nC + m 2 H 2   (16)EndothermicCarbonization
Notes: coefficients of soot and gaseous products are represented by n1, n2, n3, n4, n5, n6, and n7. While w shows the quantity of water per kmol of feedstock used, a represents a quantity of oxygen per kmol of feedstock.
Table 3. Main components of biomass. Reprinted with permission from ref. [26]. Copyright 2013 Elsevier.
Table 3. Main components of biomass. Reprinted with permission from ref. [26]. Copyright 2013 Elsevier.
BiomassLignin (%)Hemicellulose (%)Cellulose (%)Others (%)
Larch plant35272612
Willow plant2519506
Coniferous plant3026422
Deciduous plant22254112
Almond shell272725n.d.
Coconut shell352524n.d.
Sunflower seed hull271827n.d.
Spruce wood282141n.d.
Birchwood192535n.d.
Wood251842n.d.
Oakwood281935n.d.
Bagasse2039383
Rice straw12253033
Wheat straw17284015
Hardwood2035387
Softwood2824417
n.d. = no data.
Table 5. Gasifying agents and their impact on syngas composition. Reprinted with permission from ref. [54]. Copyright 1999 Elsevier.
Table 5. Gasifying agents and their impact on syngas composition. Reprinted with permission from ref. [54]. Copyright 1999 Elsevier.
CharacteristicAirOxygenSteam
FeedstockMSWMSWMSW
Temperature (°C) 777800900
ER0.40.2-
Moisture Content (%)7.598.31-
Catalyst No CatNo CatNo Cat
S/B --0.8
Char Yield (wt. %)-15.57.9
Tar Yield (wt. %)11.4 (g/m3)43.50.2
LHV (MJ/Nm3)2.48.515
CO2 (vol%) 1535.517.5
H2 (vol%) 511.828
CO (vol%) 1930.316.5
CH4 (vol%)510.321
Carbon Conversion Efficiency (%) 61-44.1
Dry Gas Yield (m3/kg) 1.4-0.5
Table 7. Operating parameters of different gasifiers using various gasifying agents and syngas characteristics.
Table 7. Operating parameters of different gasifiers using various gasifying agents and syngas characteristics.
Performance
Criterion
UnitsDFBG
(Air)
FFBG (Air)PlasmaDD
Fixed Bed
(Air)
UD Fixed
Bed (Air)
BFBG (Air)CFBG (Air)EFG (Air,
Steam)
Stratified Twin-fired Updraft
Fixed Bed (Air)
Stratified
Downdraft
Fixed Bed (Air)
Two-Stage
Fixed bed
(Air)
Gas quality—Tarmg/m320,000–40,000<50015–30010,000–150,0003000–40,0004000–20,00030<5020–200<5
Technical complexityDegreeHighHighHighSimpleSimpleMediumMediumHighMediumMediumMedium
Catalyst/bed systemTypeRecirculating supportedCharNoneCharCharFluidized bedFluidized bedcharCharCharChar
Scale including modularityMW10–10002–3<10<2010–10010–1003–100–21–101–5
Exit gas temperature°C830–850800570–880700–800400–650800–1000750–900800650<750750–800
Gasification temperature°C850–870850>3000900–10501150–1300800–1000750–900110013001000–11501100–1200
Fuel flexibility—moisture%<2534<30<20<60<55<5540–6010–15<10<50
Fuel flexibility—sizemm<1530–501–510–3002–50<5<15<50<501–535–45
Carbon conversion efficiency—CCE%9590>95<8540–8570–9080–9084–949590–9599
Cold gas efficiency—CGE%90–9375–80065–9020–6070–9050–7065–8080–9080–9080–90
Gas composition—N2vol%<5450–1465848460494030
Gas composition—CH4vol%10–13201–52–32–74–60–211–21–2
Gas composition—CO2vol%15–25155–1511–138–1011–2516–1812–16118–1015–20
Gas composition—COvol%25–352036–5210–2215–2015–2215–1825–312225–2716–20
Gas composition—H2vol%25–401845–5515–2110–1412–2615–1755–581722–2330–32
Gas quality—LHVMJ/m313–205–610–254–65–64–74–610–1255–66–7
References [128,137,138,139][140][141][128,137,140,142,143,144,145,146,147][128,137,140,142,143,144,145,146,147][128,137,142,143,148][143,149][137,150,151][152][128,153,154][128,155,156,157,158]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Maitlo, G.; Ali, I.; Mangi, K.H.; Ali, S.; Maitlo, H.A.; Unar, I.N.; Pirzada, A.M. Thermochemical Conversion of Biomass for Syngas Production: Current Status and Future Trends. Sustainability 2022, 14, 2596. https://doi.org/10.3390/su14052596

AMA Style

Maitlo G, Ali I, Mangi KH, Ali S, Maitlo HA, Unar IN, Pirzada AM. Thermochemical Conversion of Biomass for Syngas Production: Current Status and Future Trends. Sustainability. 2022; 14(5):2596. https://doi.org/10.3390/su14052596

Chicago/Turabian Style

Maitlo, Ghulamullah, Imran Ali, Kashif Hussain Mangi, Safdar Ali, Hubdar Ali Maitlo, Imran Nazir Unar, and Abdul Majeed Pirzada. 2022. "Thermochemical Conversion of Biomass for Syngas Production: Current Status and Future Trends" Sustainability 14, no. 5: 2596. https://doi.org/10.3390/su14052596

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