Sonu Kumar*,
Rupesh Palange and
Cataldo De Blasio
Laboratory of Energy Technology, Faculty of Science and Engineering, Åbo Akademi University, Rantakatu 2, 65100 Vaasa, Finland. E-mail: sonu.kumar@abo.fi; Tel: +358 466376284
First published on 22nd July 2025
The paper presents a comprehensive overview of the latest biomass gasification technologies discussing various configurations like fixed bed, entrained-flow, and fluidized-bed gasifiers along with advanced systems like plasma, multistage biomass, supercritical water and solar gasifiers. The article gives a comparative analysis by examining the performance levels, operational efficiency, and technical parameters within each gasification system. Additionally, the review delves into various gasification modeling approaches which include thermodynamic equilibrium models, kinetic models, computational fluid dynamics and data-driven models using machine learning. These modeling techniques are assessed in terms of their governing equations, solution methods, and accuracy in predicting gasification outcomes. The study also investigates the conversion of syngas and power into green fuels such as methane, hydrogen, and ammonia, with a focus on their techno-economic feasibility and sustainability through life cycle assessment. The review highlights multiple synthesis pathways for green fuel production and evaluates the most efficient methods for maximizing output while maintaining cost-effectiveness. Key research gaps and challenges within gasification technology, modeling methodologies, and green fuel synthesis are identified, emphasizing the need for improvements in gasifier operational efficiency, tar reduction, and model precision.
Mathematical modeling plays a crucial role in design and analysis of industrial and pilot gasification systems. Popular modeling approaches include thermodynamic equilibrium models, kinetic models, and artificial neural network (ANN) models.26 A majority of biomass gasification investigations depend on equilibrium models with these models accounting for close to 60% of reported studies. Analysis using CFD or ANN methods remains minimal throughout the research.27 System analysis through CFD modeling uses primarily mass, heat and momentum conservation principles to provide powerful insights into fluid dynamics and heat transfer besides examining energy flows and chemical reactions and mass and momentum conversions within a system.24 Approximately 72.5% of studies related to gasification modeling have utilized the non-stoichiometric approach in equilibrium model.28 Numerical solutions of gasification for multiphase flows can be categorized into Eulerian formulations, for instance the two-fluid, and Lagrangian formulations including the discrete element method and the multiphase particle-in – cell method.29 Also, the data driven model based on ANN proved very accurate for predicting syngas production and chemical breakdown throughout the fixed-bed gasifier biomass conversion process.30
Gasification using air generates a fuel gas with a low heating value, typically ranging from 4 to 7 MJ Nm−3.31 When O2 and steam are used in coal gasification, the resulting synthesis gas has a heating value ranging between 10 and 18 MJ Nm−3.32 Despite extensive research and numerous publications on gasification, large-scale commercialization of these technologies remains limited. The highest reported gasification efficiencies for various feedstocks are: coal at 68.5%, pine needles at 76.0%, plywood at 76.5%, and lignite at 74.0%. Additionally, cold gasification efficiency typically ranges between 63% and 66%.33
This review explores recent advancements in gasification technology, integrating both traditional and advanced modeling approaches to predict gas composition and assessing the sustainability of green fuel production. It critically evaluates various gasification technologies by analyzing their cold gas efficiency (CGE) and carbon conversion efficiency and also offering comparative insights into different modeling techniques. The study also examines cost-effective biomass-to-energy conversion pathways specializing in hydrogen (H2), methane (CH4), and ammonia (NH3) production. A comprehensive techno-economic assessment is presented, considering factors such as total production cost, capital investment, transportation expenses, energy costs, payback periods, and the carbon-negative potential of green fuel synthesis through life cycle analysis. Additionally, the study investigates the integrated synthesis of ammonia and methanol within gasification systems, identifying operational synergies to enhance process efficiency. By highlighting current research gaps in gasification technology and modeling approaches, this review provides a roadmap for advancing sustainable energy solutions and promoting innovation in green fuel production.
The two types of fluidized bed gasifiers include bubbling and circulating which differ based on their flow characteristics and heat exchange capabilities. A bubbling fluidized bed uses sand or alumina particles to form an inert condition before the gas velocity reaches the minimum fluidization level. Higher gas velocities in circulating fluidized beds create conditions which cause particles to exit the system with the gas stream. The gasifier receives these particles in a cyclone device which acts to return them to the gasifier system.36
The solid particles in a cylindrical fluidized bed stay suspended through gas or liquid flow to achieve optimized surface contact for efficient reactions. Excellent heat and material transfer functions together with uniform temperature control alongside the ability to handle various fuel types represent the main advantages of this system. Dust production together with particulate formation represents a significant downside because these pollutants affect equipment in the subsequent stages.35 Fluidized bed gasifiers deliver several benefits through their perfect temperature control system with excellent heat exchange capabilities and efficient hot spot elimination abilities. The large fuel inventory alongside good gas–solid mixing produces devices with easy operation as well as high reliability. The gasifiers exhibit multiple advantageous characteristics because they achieve high specific capacity by delivering quick reaction rates while maintaining short solid residence times and working at partial load levels between 40–120%. These systems need minimal physical space while accepting various fuel properties and particles together with strong carbon conversion for initial gas filter operations. The product gas obtained from these units contains minimal phenolic substances while requiring lower overall investment costs.35,37
The challenges of these gasifiers include dual temperature requirements for different feedstocks together with more complex operations than fixed beds and substantial pressure loss throughout the system and restricted capacity caused by gas velocity effects. Non-molten ash along with elevated energy requirements and abundant dust matter in the gas phase represent different disadvantages of these units.
The studies on existing commercial and near-commercial biomass gasification technologies indicate that directly heated fluidized bed (FB) gasifiers stand as the most frequently implemented systems as illustrated in Fig. 3.38 The entrained flow gasifier utilizes oxygen as its gasifying element at temperatures between 1200–1500 °C while maintaining extremely brief residence times of a few seconds. High operating temperatures lead to reduced tar formation and combustible gas concentration while improving the conversion of char materials. The processing capabilities of entrained flow gasifiers now hold greater significance because they combine quick processing and efficient small particle handling with maximum carbon conversion rate. The combination of high capacity with short retention period and strong thermodynamic efficiency makes engineers choose this technology as their primary solution.39,40 Table 1 has discussed the important differences between the three types of gasifiers and mentioned its limitations also.
Parameters | Fixed bed gasifier | Fluidized bed gasifier | Entrained flow gasifier |
---|---|---|---|
Temperature and ash handling | Non-homogeneous temperature distribution, temperature ranges above ash melting point | Homogenous temperature distribution ranging between 800 to 1000 °C. Operating in conditions below the ash-softening point ensures they prevent agglomeration | High operation temperatures reaching 1400 °C while surpassing the ash melting point thus producing molten slag |
Particle size and residence time | It uses large size feedstock particles and has high residence time | It accepts wide range of particle size and has less residence time compared to fixed bed | It requires a dry and uniform size biomass particle and has very less residence time |
Mixing and reaction characteristics | There is limited mixing between solid and gas | This gasifier allows an excellent mixing of solid and gas phase | This gasifier allows simultaneous mixing and co-current flow of biomass and gasifying agent. Has high carbon conversion efficiency |
Capacity and scalability | This is good for small or medium scale production | It is good for industrial or large level of production and has high gasification intensity | Since it operates at very high pressure and temperature, so it is suitable for large scale application |
Product gas characteristics | It produces low heating value gas with small amounts of tar | It produces low tar with stable composition and produce gas with high fly ash content | Its yield quality is high, and has tar free syngas with less methane content |
Fig. 4 describes the detailed working process of the plasma gasifier where it uses plasma torch for heating the feedstock particles to convert into syngas and byproduct like ash. Plasma gasification operates as an innovative technology which utilizes electrically ionized gas that reaches 10000 °C temperatures produced by plasma torches operating under pressures ranging from 1 to 30 bars.46 The feedstock enters through the reactor's top section while the gasifying agent is delivered to its sides to start the required reactions. Plasma gasifiers utilize thermal plasmatic systems as their primary mechanism for high-efficiency heat generation.47 Plasmas inside the reactor establish a hot electrically conductive column which generates incredibly high temperatures. The process completely disintegrates all materials which then transform into gaseous products.48 The system efficiently deteriorates diverse biomass feedstocks directly through its process without needing preprocessing techniques. The process delivers outstanding decomposition capacity that produces substantial energy output. Its operation produces minimal emissions together with short residence times that result in improved overall environmental advantages and operational performance.49,50 The high efficiency of plasma gasification requires substantial investments and operation costs. To power the reactor of a 3600 MW plant we need 115 kW h per ton of MSW excluding other system elements. The technology proves commercially worthwhile since research demonstrates that plasma gasification of MSW generates approximately $3.35 million in annual profit. Through plasma gasification we achieve waste management efficiency and create economic and environmental benefits that enable a circular economy operation.49,51 Therefore, plasma gasification demonstrates a substantial environmental advantage through its process which produces around −31 kg CO2eq in negative greenhouse gas emissions. The environmental impact of plasma gasification remains low even when accounting for the small emissions of particulate matter and heavy metals. A long-term waste management and energy production solution becomes possible through plasma gasification because the investment can recover costs within 18 years despite high installation and management expenses.45,52
![]() | ||
Fig. 5 Line diagram of multistage gasifier.53 |
Through this technology operational efficiency increases while the final syngas product contains decreased amounts of tar. The multiple operating components create substantial operational complexity since their continuous coordination procedures must be maintained. The primary objective in creating multistage gasifiers was developing gas that contained the lowest possible tar quantities. The review demonstrated that temperature levels in the multistage heating and gradient chain gasifier (MHGCG) were highest during the drying phase compared to the subsequent oxidizing and reducing stages. Enhancing drying stage heating while minimizing oxidizing and reducing stage temperatures led to an important elevation of gasification efficiency over standard operational parameters. The gasification process achieved 56% efficiency with a biomass weight loss of 43% when the oxidizing stage equivalence ratio was set to 0.28. Provided gas composition consisted of 4% O2, 10% CO2, 14% H2 and 24% CO and the final product concentrations measured at NO: 0.024%, NOx: 0.025% and SO2: 0.032%.54 A fluidized bed reactor consisting of multiple stages is simulated through Aspen Plus for Prosopis Juli flora air gasification during its pyrolysis, combustion and reduction phases. The semi-detailed kinetic approach uses both reaction rates and hydrodynamic principles for optimized gasifier performance assessment. The computer model indicates gasifier temperature together with air-to-biomass ratio serves as main variables which impact CO gas output. To achieve maximum gas calorific value the system requires a gasifier structure with a 2 m height and 0.5 m diameter and an equilibrium ratio of 0.24 to achieve 65% cold gas efficiency. The reactor technology generates valuable product gas containing high energy value.56 Saleh et al.57 conducted a study of multistage air inlet modifications to downdraft gasifiers enhanced gas quality through higher heating values and lower tar levels. The evaluation of performance operated through three metrics: equivalent ratio (ER) and preheated air temperature and air ratio. The peak performance of the system occurred at an equivalent ratio of 0.4 at 902 °C leading to optimized CO and H2 production rates and preheating the air resulted in increased LHV from 5254 to 5976 kJ kg−1. With the optimal ratio of 40:
60 the system successfully diminished tar content from 50.02 to 27.82 mg Nm−3. Implementation of a multi-stage gasifier operated through variable temperature steam gasification resulted in a 15-fold CO production increase during the second stage with Gasification Carbon (G-C) added substances and produced 2.32 times more total syngas. The addition of G-C enabled researchers to control the H2/CO ratio thereby benefiting subsequent synthesis operations. The process achieved both lower energy consumption and reduced CO2 emissions and performed efficient CO2 capture and conversion. The method improved the syngas yields for cleaner energy applications.58 Therefore, the multistage biomass gasification process decreases syngas tar levels to a point where it becomes usable by internal combustion engines alongside gas turbines without needing expensive deterring equipment. Such gasification systems create cleaner gas outputs than single-stage gasifiers though they become less efficient for wood biomass processing. The technology demands superior quality feedstock which reduces its possible uses. Technology operates against entrained-flow gasifiers when evaluating gas purity.
The 1 ton per h treatment capacity simultaneous saccharification and fermentation (SSF-SCWG) model was used for studying how temperature and reactant concentration influence SCWG product yields.66 Energy levels increased along with temperature while mass concentration levels decreased the production efficiency of hydrogen and gasification. The combination of maximum hydrogen production reached 0.139 kg kg−1 accompanied by 108.832 mol kg−1 gasification yield.67 The SCWG also allows the production of biofuels and valuable products from Kraft Black Liquor (KBL).68 Through using a high-heating-rate batch reactor in sewage sludge gasification SCWG operations the process generated considerable higher levels of hydrogen compared to conventional production methods. The reaction temperature influenced all parameters positively and reached their maximum levels of 20.66 mol kg−1 hydrogen yield at 750 °C and both without and with a catalyst. The combination of steam reforming with water–gas shift and pyrolysis together produced H2 and CO2 from 550 to 750 °C with the water–gas shift taking the lead role when reaction time was extended. Higher reaction temperatures combined with prolonged residency duration minimized the production of CO and CH4.69 The research utilizes municipal waste leachate as its focus for creation of Synthetic Natural Gas (SNG) through catalytic improvements using Nickel-based catalytic conversion. The process takes place during the following operational stage beyond the SCWG reactor. The produced syngas consisted of hydrogen and methane gases within a concentration range of 25–47 vol% H2 and 11–18 vol% CH4.70
A study developed thermodynamic equilibrium model for pig manure SCWG then investigated how different operating conditions affected heat generation together with system performance.73 The results demonstrated how lowering preheating water temperature while increasing the slurry concentration together with higher gasification temperatures improved both heat production and system efficiency up to 95.53% using 1:
1 water-to-slurry ratio and 94.90% with 70 wt% slurry concentration. System efficiency reached 39.8% based on exergy analysis and most exergy losses took place in the heat exchange and oxidation stages. Therefore, using SCWG provides an environment-friendly and optimized process to transform coal into hydrogen products compared to conventional gasification systems. High solubility along with reactivity and diffusivity properties in supercritical water allow coal to transform into hydrogen-rich gases at manageable temperatures. At the reactor's bottom natural sedimentation of inorganic salts involving N, S and Hg takes place. Filtration of H2 from CO2 occurs effectively through critical point pressure control without the need for specialized separation equipment.71–73
Fig. 7 shows the detailed operating system of solar gasifiers and describes the utilization of solar thermal energy for the generation of syngas. Solar thermal systems apply focusing mirrors to collect sunlight which produces high-temperature heat exceeding 1273 K. These systems apply optical surfaces to achieve high-temp focal points which operate with efficient heat delivery. The cavity receiver design reduces heat losses to less than 30% which boosts system performance. Solar reactors employ two operational methods to heat their particulate solid feedstock through directly irradiated and indirectly irradiated reactors.76,77 The eqn (1) and (2) can be helpful to find the energy and thermal energy conversion efficiency as mentioned below.
![]() | (1) |
![]() | (2) |
Taylor et al.78 constructed a 2 kW solar furnace to operate the first fluidized bed reactor by inserting a silica-glass tube vertically for solar gasification research. The process of fluidizing coconut charcoal using CO2 flows operated at 2–15 m min−1 resulted in a 10% energy conversion efficiency but lost significant power due to heat escaping through radiation and conduction and gas escape. The packed-bed reactor operated at 40% efficiency because of its performance improvement over other configurations. A study combines a solid oxide fuel cell (SOFC) which runs on hydrogen–carbon monoxide syngas with a coal-fired combined cycle and concentrated solar energy. The combination of energy systems yielded enhanced performance characteristics by reaching energy efficiency ranges from 70.6% to 72.7% and exergy efficiency rates between 35.5% and 43.8%. The utilization of different coal types resulted in CO2 emission levels between 18.31 kg s−1 while using 10 kg s−1 fuel.79 Moreover, the rise in power and chemical sector need for solid carbonaceous feedstocks including coal and biomass with waste materials drives fast development of gasification technologies. The alternative solar-driven gasification system corners process heat from concentrated solar radiation to supply high-temperature requirements. The process both improves gas synthesis efficiency and cuts down CO2 emission rates through solar-energy-driven calorific value enhancements of raw materials. The elimination of air separation units increases economic viability and makes solar gasification an efficient technique for storing solar energy through mobile chemical compounds.80
Additionally, it can be seen that the gasification technologies are shifting towards the green energy operated plant, and more research has been done in the field of solar based energy system (See Fig. 8). Since the electricity cost has a huge impact on the total investment cost of the gasification system. Therefore, making gasification system more economical and reducing the carbon emission there is necessity to explore the green source of power supply.
Name | Chemical reaction | Heat energy (MJ kmol−1) |
---|---|---|
Boudouard reaction | C + CO2 ↔ 2CO (R.1) | +172 |
Water gas or steam | C + H2O ↔ CO + H2 (R.2) | +131 |
Methane reaction | C + 2H2 ↔ CH4 (R.3) | −74.8 |
Oxidation reaction | C + O2 → CO2 (R.4) | −394 |
CO + 0.5O2 → CO2 (R.5) | −284 | |
CH4 + 2O2 ↔ CO2 + 2H2O (R.6) | −803 | |
H2 + 0.5O2 → H2O (R.7) | −242 | |
Shift reaction | CO + H2O ↔ CO2 + H2 (R.8) | −41.2 |
Methanation | 2CO + 2H2 → CH4 + CO2 (R.9) | −247 |
CO + 3H2 ↔ CH4 + H2O (R.10) | −206 | |
CO2 + 4H2 → CH4 + 2H2O (R.11) | −165 | |
Steam reforming reactions | CH4 + 0.5O2 ↔ CO + 2H2 (R.12) | −36 |
CH4 + H2O → CO + 3H2 (R.13) | +206 |
![]() | (R.14) |
![]() | (3) |
![]() | (4) |
![]() | (5) |
![]() | (6) |
![]() | (7) |
![]() | (8) |
Another approach for thermodynamic equilibrium modeling of gasification process is the non-stoichiometric method which is based of minimization of Gibbs free energy. It also includes the calculation of chemical potential, and Lagrange function and the related eqn (9)–(13) are given below.
![]() | (9) |
μi = ΔG0f,i + RT![]() ![]() | (10) |
![]() | (11) |
![]() | (12) |
![]() | (13) |
The total Gibbs energy can be calculated using eqn (9) (called Gibbs–Duem equation) where ni and μi represent the number of moles and chemical potential of species.88 This modeling approach is based on the principle that, at equilibrium, the total Gibbs free energy of the system reaches a minimum under specified temperature and pressure conditions. This method does not require a predefined reaction pathway, instead, it considers all possible species and phases and determines their equilibrium composition by minimizing the system's Gibbs free energy, subject to elemental mass balance constraints. As a result, it offers a robust and comprehensive framework for predicting the composition of the syngas in complex gasification systems.
The accuracy of both stoichiometric and non-stoichiometric equilibrium models in biomass gasification can be significantly enhanced by incorporating actual reaction conditions observed during operation. Various modified equilibrium models have been developed, which consider factors such as equilibrium constants, experimentally determined gas compositions, specific reaction stages, operating parameters, and empirical correlations. These refinements enable more realistic predictions of product distribution and process performance under practical gasification conditions. Mountouris et al.89 developed a model based on equilibrium constants used water–gas shift and steam reforming reactions with partial mass balance system for carbon, hydrogen, oxygen alongside heat balance. The model accounts for solid carbon formation and provides exergy data necessary for optimizing system processes. Silva and Rouboa90 presented a two-stage equilibrium model that establishes equilibrium composition in stage one and stage two operations with gases only while omitting solid carbon. The equilibrium constant was corrected through multiplicative factor to improve equilibrium modeling. Additionally, Aydin et al.91 developed a semi-empirical equilibrium model that integrates two correction variables to enhance the equilibrium constant of the methanation and water–gas shift reactions. The corrections which rely on gasification temperature, equilibrium temperature and equivalence ratio (ER) measurements resulted from comparing theoretical models to downdraft woody biomass experimental data through the Levenberg–Marquardt method. The modified model shows improved ability to predict concentrations of gaseous species along with tar yield in producer gas. This modified mathematical framework can produce a better estimation of both producer gas composition and obtained tar content.92 Table 3 summarizes different studies which implemented TEqM for simulation of gasification process along with important findings and limitations.
Reference | Feedstock | Gasification agent | Remark | Future research possibilities |
---|---|---|---|---|
93 | Rubber wood | Air and water vapour | Studies demonstrated that water vapor functioned as a gasifying agent to produce superior-quality syngas than when using air. The experiment utilizing air generated a hydrogen yield at 35% by volume but switching to steam as the gasifying agent increased the yield to 65% by volume | The gasifying agents are very limited which need to be explored for getting better conversion of biomass to syngas |
30 | Almond shells | Air, pure oxygen, steam | The research proves that updraft gasifiers can produce syngas which contain higher hydrogen content than other gasifier designs because of its efficiency. The difference between predicted results and experimental findings remains below 10% on average | The primary disadvantage of the updraft gasifier is the high tar content in the output gas, which reduces its lower heating value |
94 | Pine sawdust | Air | The study revealed that increasing the H2O content consistently enhanced the yield of effective gases (CO and H2), whereas an excess of CO2 inhibited their production | The removal of carbon completely proved difficult to achieve during entrained-flow gasification under analyzed conditions. The study requires more investigation to establish methods which can both decrease and eliminate carbon formation under these conditions |
95 | Municipal solid waste | Air | The bioenergy system operated at a production capacity of 3.92 MW electrical power together with 608.8 m3 h−1 hydrogen production when fed with 1.155 kg s−1 biomass. The established design parameters resulted in an energy utilization factor of 34.71% together with a total exergy efficiency of 29.44% and overall exergy destruction rate of 11![]() |
The combustion chamber and gasifier experienced maximum exergy destruction because intense chemical processes occurred within their structure. Boosting system exergy efficiency would result from lowering these exergy destruction rates |
96 | Sewage sludge | O2 and steam | The model has been adapted to include the sulfur content in sludge, along with char conversion and tar formation. For the temperature range of 900 to 1150 K, the average mole fractions of the primary gas components were determined to be 9.76% for H2, 11.80% for CO, 9.84% for CO2, and 2.97% for CH4. As the temperature increases, the molar concentrations of H2 and CO rise, while those of CO2 and CH4 decline | The AAE, SSE, and RMSE values for all product gas components, PGY, LHVp, CGE, and CCE were found to be below 10%. These errors could be further reduced by utilizing an alternative solving model |
97 | Polypropylene, polyethylene terephthalate, biomass (straw) | Air | An increase of plastic content in gasification feedstock reaches its maximum heating value at 5.78 MJ Nm−3 alongside its highest tar concentration of 72.89 g Nm−3 | Since the tar content has risen more which reduces the syngas yield which is unwanted. So, this can be reduced further by taking different composition of feedstock |
98 | Woods & sewage sludge | Air, pure oxygen, steam | This study evaluates syngas composition, tar and char yields, gasification temperature, cold gas efficiency, and the lower heating value for different biomass feedstocks characterized by specific ultimate analysis. The predictions are conducted for varying equivalence ratios and moisture contents | This model determines methane concentration levels without needing correction factors to make predictions which is a strength that equilibrium models often lack |
99 | Pine kernel shells (PKS) | Air, steam | The stoichiometric and non-stoichiometric models received validation through experimental data obtained from a semi-pilot scale bubbling fluidized bed gasifier operating with pine kernel shells (PKS) feedstock. The stoichiometric model proved to be more accurate than the non-stoichiometric model for predicting gas composition along with gasification efficiency | The prediction of CO and methane at a stoichiometric ratio (<0.2) still shows notable deviations from the experimental data, which can be further minimized |
100 | Waste tires | Steam | The study identified the optimal conditions for maximizing H2 yield in the supercritical temperature range as 599.8 °C, 23.2 MPa, and 5.5 wt%. Conversely, the ideal parameters for achieving maximum CH4 yield in the supercritical temperature range were 551 °C, 27.2 MPa, and 18.3 wt%. For the transition temperature range, the optimal conditions for CH4 yield were determined to be 380 °C, 26.4 MPa, and 7.9 wt% | The predicted H2 yield differences from experimental results reached significant levels. Percentage errors amounted to 46.5%, 27.1% and 48.3% |
101 | Napier grass | Air | The research demonstrates that for assigned temperature ranges and ER values the model achieves RMS results of 0.0227 and 0.1108 which verify the precise simulation of gasification behavior | The updated model predicted elevated levels of H2, and CO2 formation with around 16% deviation and ash generation but it calculated reduced concentrations of CO and CH4 |
102 | Coal | Air and steam | An optimization process using the Taguchi method and utility concept along with TEqM to maximize syngas calorific value and minimize CO2 yield during coal gasification. The Taguchi method identified two optimal control variable sets, achieving a calorific value of 3.59 MJ m−3 and CO2 yield of 6.25%. The utility concept was used to simultaneously optimize both objectives, with air supply, steam supply, and H/C ratio of coal being the most influential parameters. Results from the utility concept showed a 3.34% and 2.30% difference in calorific value and CO2 yield compared to Taguchi's findings | The optimized model has assumed ideal gas behavior and does not consider the formation of tar which is an important byproduct formed in the gasification process |
The equilibrium models are still widely used because of their simplicity and computational efficiency. These models are especially valuable for predicting the theoretical maximum syngas yield and establishing thermodynamic feasibility boundaries for gasification processes. Improving equilibrium models is essential for enhancing the accuracy and reliability of biomass gasification simulations, particularly under practical, non-ideal conditions. Bijesh et al.96 advanced a stoichiometric thermodynamic model of equilibrium of sewage sludge gasification that accounts for sulfur content, char transformation, and four classes of tar compounds. The model showed strong agreement with experimental data, achieving R2 values above 0.90 and p-values below 0.05. The results demonstrated that hydrogen (H2) and carbon monoxide (CO) concentrations increased with temperature, while carbon dioxide (CO2), hydrogen sulfide (H2S), and tar concentrations decreased. In another example, a recent review by Carine et al.103 demonstrates how these models are often applied to carry out fast screening of feedstocks and operating conditions. However, a fundamental limitation of these models is their assumption of complete chemical equilibrium, which overlooks reaction kinetics, intermediate species, tar formation, and char reactivity. As shown by Ahmed et al.,104 there is always an over-estimation of the hydrogen concentration and underestimation of tar or carbon residue, especially in multistage and low temperature gasifiers by using equilibrium models. After a thorough review, it can be confirmed that fact that although these models are statistically valid they may fail to reflect the dynamic behavior of real systems. To address these problems, hybrid equilibrium approaches are currently being developed in which the cores principles of equilibrium model are combined with either empirical or semi-empirical corrections to address devolatilization, and char-gas effects.
There are basically two modeling methods in kinetic models which are semi empirical kinetic and comprehensive kinetic model.28 Semi-empirical kinetic models establish local equilibrium in specific reactions and gasifier regions through their calculations of both kinetic-controlled concentrations and temperatures in other areas. The models provide suitable computational efficiency together with accuracy benefits that allow researchers to understand complex processes in biomass and coal gasification. Experimental data including temperature-dependent reaction rates and feedstock behaviors allows the models to produce accurate predictions regarding the gasification efficiency as well as char reactivity and syngas composition. Researchers utilize these tools extensively for hydrogen optimization as well as assessment of low-temperature char reactions and reactor phase simulation.109 The comprehensive kinetic modeling methods account for both volatile and char reactions rates by tracking temperature and species changes during reactor progression through time. The semi empirical models demonstrate higher accuracy compared to comprehensive models when the gas phase achieves equilibrium conditions because they need fewer reaction rate laws and parameters.110
Empirical models make reaction kinetics simpler to handle by using polynomial approximations, Arrhenius-type expressions, or power-law equations. A research study developed a flexible polynomial model for following rate changes during CO2 biomass char gasification as conversion increases.111 The model received verification through 24 TGA experiments where both modulated and constant reaction rate (CRR) temperature programs were utilized to eliminate thermal deactivation and measurement errors. The model parameters were refined and optimized using the least square method and when compared with experimental data the results demonstrated high reliability. These models excel at determining the influence and structural complexity of minerals and catalytic effects present in biomass chars.111,112 Other modeling methods include single step, multiple steps, Arrhenius based model, and Langmuir–Hinshelwood model. The Global or single-step kinetic models convert gasification into a single reaction for quick estimations without accounting for intermediate reaction steps. Whereas having separate stages in multi-step kinetic models that include pyrolysis and char oxidation with additional gas-phase reactions results in better predictions regarding gas composition behavior.113 Another approach is Arrhenius-based models, which have gained widespread use to explain reaction rates through their dependency on activation energy and temperature use of the Arrhenius equation. The prediction of char gasification rates proves most effective when using these models.114 Similarly, Langmuir–Hinshelwood models provide a useful tool for catalytic gasification processes because they help to explain how gas–solid reactions are influenced by surface coverage and adsorption phenomena. Eqn (14) and (15) show the correlation between the kinetic reaction rate, reaction component concentration and reaction equilibrium constant for the chemical reaction involved in gasification process.34,115
The Kinetic model basically works on kinetic reaction rate:
![]() | (14) |
Arrhenius equation for solving kr:
![]() | (15) |
Different studies on gasification based on kinetics models along with their limitations and future potential are summarized below in Table 4.
Reference | Feedstock | Gasification agent | Remark | Future research possibilities |
---|---|---|---|---|
116 | Corn stover | Air | The study demonstrates that optimum syngas production could be achieved with gasification settings of 850 °C temperature, an O/C ratio of 0.45 and H2O/C ratio of 0.6. The experimental conditions produced 0.77 m3 kg−1 of syngas with 66.2% cold gas efficiency. The current results indicated a heat output power measurement of 15.42% and a heat efficiency rate of 60.99% | This paper introduces a model which matches experimental observations better than the thermodynamic equilibrium model does regarding accuracy. The total exergy efficiency level stands at 41.64% but still has potential for development. The system's overall performance could be optimized through improved methods of energy utilization efficiency |
117 | Pomegranate wood | Air | A power output of 7 kW generated syngas with higher molar ratios of H2 at 21.5% and CO at 30.08%. The power load increase from 7 kW to 10 kW led to H2 concentrations decreasing by 13.35% and CO concentrations eroding by 13.76%. The model generated syngas composition results with suitable precision through mean absolute percentage errors reaching 8.91% for H2 and 1.98% for CO which satisfied the researchers | The CCE and CGE in the model are increasing as electric load of gasifier is increasing so further higher power can be investigated to analysis the syngas quality |
118 | Sewage sludge | Air/steam | A sensitivity analysis of the proposed hybrid SSG model demonstrated that the syngas compositions together with cold gas efficiency (CCE) are strongly affected by changes in flow rates of both gasifying agents (air and steam) and sewage sludge feedstock and by operational temperature and pressure | The validation approach in this research relies exclusively on experimental measurements obtained from fluidized-bed gasifiers which can be done on different gasifier |
119 | Sawdust | Air/steam | This study reveals that the dry gas yield increased by 8.1% as the steam-to-biomass ratio rose from 0.61 to 2.7, while the tar yield declined by 7.25%. Additionally, as the temperature increased, the gas yield (DGY) consistently grew from 1.72 to 2.0 Nm3 kg−1 | Since for engineering applications, the tar yield should be below 0.5 g Nm−3. However, in this study, the lowest recorded tar yield was 8.45 g Nm−3. To achieve a significant reduction in tar yield, further investigation into the use of various catalysts is required |
120 | Almond shells and hazelnut shells | Air and steam | The kinetic model proves a better choice for biomass gasification modeling since it requires parameters that are difficult to satisfy within the thermodynamic equilibrium approach. The current model provides a maximum relative error of 14.6% for CH4 prediction when used for air gasification and achieves 12.8% maximum relative error during air and steam gasification | The prediction of CO yielded the greatest deviation from experimental data due to an error of 20% which resulted in overestimated values. Absolute errors from maximum CO2 prediction results reached 29.3% and 26.4% respectively. The model's performance requires additional refinement since current deviations appear unacceptable |
121 | Lignocellulosic | Steam | Under the tested reaction conditions, residence time had little influence on gas formation during SCWG of both biomass model compounds and real biomass. At temperatures between 450 and 550 °C, the SCWG process for different model compounds and real biomass was almost entirely completed within 10 minutes | In developing the general kinetic model for predicting gas yields from real biomass during SCWG, it was assumed that no interactions occur among the three model components. However, interactions may influence the gasification process when mixtures of different components are involved |
101 | Wood residue | Air/steam | The optimization model reveals that the highest syngas yield of 78.6 vol% is achieved at a temperature of 900 °C, an ER of 0.23, an S/B ratio of 0.21, and a moisture content of 30 wt% | This study did not account for tar formation and overlooked the non-uniform temperature distribution within the gasifier |
122 | Napier grass | Air | The study shows that the highest syngas yield and higher heating value (HHV) achieved were 69.42 wt% and 8.14 MJ Nm−3, respectively, under optimal conditions of 850 °C, an equivalence ratio (ER) of 0.3021, and a moisture content of 15.69 wt%. The syngas yield, HHV, carbon conversion efficiency (CCE), and cold gas efficiency (CGE) were reported as 82.51% and 30.69%. Furthermore, the average RMSE values for process temperature and ER were determined to be 0.025 and 0.033, respectively | The more efficient kinetic model requires evaluation for its performance with different biomass types through future research. It must analyze optimal conditions from an economic standpoint because different cost-effective optimizations exist but have not been determined yet |
123 | Pet coke | Steam and CO2 | The paper concluded that CO2 plays a minimal role in the process and as demonstrated in the following section, the reaction involving CO2 can be disregarded if the temperature does not substantially exceed 1000 °C | In the research paper, it illustrates that at a temperature of 1000 °C with X(H2O) at 0.4, there is a greater discrepancy between the predicted and experimental data, which can be further minimized |
124 | Pine pellets and chips | Air | They show that the maximum absolute error for H2 prediction was just 4.4%. Additionally, the predicted tar concentration ranged from 20 to 42 g Nm−3 and decreased with increasing equivalence ratio, temperature, and biomass particle size | The paper demonstrates the least precise prediction of H2 gas which shows room for enhancement in accuracy levels |
125 | Rubber wood and refused devised fuel | Steam | A three-zone kinetic model for evaluating rubber wood and refused derived fuel (RDF) gasification by studying five reduction zone reactions. Taguchi optimization showed that methanation represented the most influential process for RDF because it accounted for 65% of calorific value and 71% of exergy efficiency. In rubber wood gasification the Boudouard reaction accounted for 49% caloric value production and the water–gas reaction directly affected exergy efficiency by 46% | The model has assumption of ideal gas behavior with neglection the formation of the tar as byproduct in the gasification process |
Kinetic models aim to represent the actual reaction kinetics involved in biomass gasification by explicitly modeling the rates of individual processes such as pyrolysis, oxidation, and reforming reactions. These models typically employ Arrhenius-type expressions to describe the temperature-dependent behavior of chemical reactions, allowing for a more detailed understanding of reaction mechanisms, particularly in relation to time-resolved phenomena such as tar cracking, pollutant formation, and intermediate species evolution. These models are especially valuable for analyzing the transient behavior of gasification systems and for designing processes where dynamic control is essential. They frequently, however, necessitate large pools of kinetic parameter data that are quite exclusive to the type of feedstock, the particle size and the type of reactor being considered. According to Sylwia et al.,126 this kind of specificity restricts the generality of the kinetic models, and the cost of their calculations becomes excessive when they are included in massive simulations. To address these challenges, recent research has focused on the development of reduced-order and semi-global kinetic models. These simplified frameworks aim to retain essential predictive capabilities while reducing the number of required parameters and overall computational cost. Such models are increasingly being used in real-time optimization, control, and system integration applications where full-scale kinetic modeling is impractical. Another example of advancements in kinetic models can be found in case of simulation of biomass staged gasification technology (BSGT), where an important issue is the accurate modeling of medium temperature devolatilization (MTD) and char gasification. Conventional kinetic models often fail to capture this process accurately due to the omission of catalytic effects particularly the influence of inherent alkali metals like potassium. To overcome this limitation, an improved kinetic approach based on the random pore model (RPM) was developed, incorporating a correction factor to account for the catalytic role of potassium in corn stalk gasification. This enhanced model, referred to as RPM+, significantly improves the accuracy of kinetic predictions and offers more reliable guidance for reactor design and simulation in BSGT systems.127
CFD modeling plays a vital role in simulating fluidized-bed and fixed-bed downdraft gasifiers. Key parameters analyzed during CFD simulations include drag force, biomass porosity, and turbulence attenuation. Due to the high permeability of the bed, a constant pressure assumption was applied throughout the reactor. Pressure drops calculations incorporated a modified Ergun equation, along with numerical solutions of transport equations using finite-rate kinetic reactions.17 Moreover, the established correlations in literature provided transport coefficients and chemical kinetics alongside the implementation of a finite volume method for accurate simulation of the gasification process. CFD software allows different gasification projects to guide optimal setup selection while performing budget-friendly evaluation of configurations and operating conditions at different scales.129
The governing equations essential for the CFD modelling of the gasification process include the generalized forms of the mass, momentum, and energy conservation equations. These equations are employed to determine the syngas composition, velocity and pressure fields, and temperature distribution within the gasifier. The CFD framework solves the conservation equations separately for the gas phase and the solid phase, acknowledging the distinct physical behaviours of each state. Given the significant interphase interactions between solid particles and the fluid medium, accurate modelling also necessitates the inclusion of the Energy Minimization Multi-Scale (EMMS) drag model in some cases. Fig. 9 depicts the data flow architecture for the CFD modelling of the gasification process. It encapsulates essential inputs such as feedstock characteristics, syngas composition, reaction kinetics, equilibrium constants, governing transport equations, discretization strategies, numerical solvers, and boundary condition specifications.
Although CFD modeling has proven effective for analyzing the gasification process, its application to commercial fluidized bed systems in combustion and gasification still requires further investigation, particularly when using the Eulerian–Eulerian Two-Fluid Model (TFM) approach. When applying Eulerian–Eulerian modeling to dense fluidized bed combustion and gasification it is difficult to achieve precise results unless assumptions spanning wide ranges of biomass particle sizes are made. A comprehensive investigation of CFD simulations analyzing bed and freeboard together is absent from current research and so are investigations using the same technique for both dense bed and riser/freeboard in commercial units. Limited CFD modeling has been used to study the tar formation process in gasifiers leading researchers to identify new investigation opportunities in this field. CFD models include multiple operation and design parameters but studying their impact on syngas production requires additional scrutiny. The shortage of comprehensive CFD simulations in biomass gasification stems from two primary reasons which include the expensive computational needs and the complex anisotropic characteristics of biomass materials. Table 5 summarizes key research studies on the application of CFD models to gasification processes, along with their major findings. This review encompasses a wide range of feedstocks, including coal, municipal solid waste, biomass pellets, Miscanthus briquettes, palm kernel shell, softwood pellets, and almond pruning. Various gasifying agents such as air, steam, nitrogen (N2), carbon dioxide (CO2), and oxygen have been considered across these studies.
Reference | Feedstock | Gasification agent | Remark | Future research possibilities |
---|---|---|---|---|
130 | Coal | Air and oxygen | The research demonstrates computational fluid dynamics (CFD) to be essential for modeling underground coal gasification processes as a research tool during system design | A discrepancy between experimental measurements and predicted results for CH4 and CO2 amounts to 16% to 30% thus suggesting further tests should be conducted to reduce this variation |
131 | Municipal solid waste | Air | The study demonstrates that the hydrogen yield is minimal at the bottom of the reactor, around 1%, but increases to a peak of approximately 23% near the top. Conversely, the CO molar fraction is highest at the reactor's bottom, reaching about 22%. The relative error for the four main syngas components stays below 12% | Mathematical models need extensive research with development efforts to improve their application scope and precision for process enhancement and plasma gasification implementation |
132 | Coal | Air and steam | Increasing the tapered angle of fluidized bed gasifier reduces the LHV and HHV of gas products but enhances the CCE, with the CGE improving from 3° to 5° and stabilizing between 5° and 11°. Higher velocities of the gasifying agent lower the LHV and HHV while boosting the CCE. Moreover, increasing the steam-to-air ratio reduces the concentrations of H2, CO, and CO2 | The experimental validation performs well yet the predicted H2 results show lower values than observational measurements. At an 11-degree angle both CGE and CCE reach maximum levels of 30% and 50% indicating areas for possible enhancement |
133 | Ecoshakti biomass pellet | Air | This research indicates that boosting equivalence ratio (ER) results in diminishing carbon monoxide (CO), hydrogen and methane concentrations together with growing concentrations of carbon dioxide (CO2) and nitrogen (N2). During the process of changing ER from 0.25 to 0.60 the air composition shifts to augment nitrogen mole percentage from 41.48% to 66.63% | Average errors from the measurements of CO gases hydrogen and CO2 concentrations reached 5.21%, 10.55% and 24.63% respectively. The model predicts CO2 values with a substantial error showing that additional improvements are needed to match actual experimental results |
134 | Miscanthus briquettes | Combination of air, steam and oxygen | According to the paper, increasing the equivalence ratio (ER) results in a decrease in the concentrations of carbon monoxide (CO), hydrogen, and methane, while the concentrations of carbon dioxide (CO2) and nitrogen (N2) increase. The presence of nitrogen in the air, along with oxygen, causes the nitrogen mole percentage to rise from 41.48% to 66.63% as the ER increases from 0.25 to 0.60 | Simulation results predict CO and H2 quantities which match experimental results well. The relative error calculations for CH4 show substantial variation while producing higher numbers from 6.5 to 64.7. The CH4 prediction accuracy requires additional simulation model enhancements for better performance |
135 | Coal | O2, steam, N2 | The research model demonstrated a carbon gas efficiency between 45% and 66.78% and carbon conversion efficiency from 40% to 50%. Laboratory tests proved that the fluidized bed gasifier with circulating mode operated better than a bubbling fluidized bed gasifier for carbon conversion performance. The highest production rate of 55% emerged when CRC 701 received an O![]() ![]() |
There is still significant deviation in yield percentage prediction of H2 and CO2 compared to experimental data for coal CRC704 |
136 | Palm kernel shell | Steam and CO2 | Higher gasification temperatures combined with S-CO2-R influence both H2 production rates as well as tar formation to a significant extent. H2 production increased by 21.4% and 20.5% during S-CO2-R operations at the higher values of 2.0 and 1173 K respectively relative to 0.4 S-CO2-R at 973 K temperature | Research shows that while the amount of tar decreases during gasification compared to the earlier devolatilization stage, the overall reduction in tar at higher reaction temperatures during both stages is not very significant |
137 | Softwood pellets | Steam | Research through the hybrid Euler–Lagrange approach of DDPM demonstrated the biomass-to-char conversion process takes approximately 40 seconds which indicates viability for laboratory-scale reactor implementation. When initialization methods implement partially converted fuel particles the calculation time decreased because it brought simulations closer to actual reactor operating points | Further research needs to investigate the complete mechanisms of tar development and breakup because tar impairment remains a fundamental barrier to fluidized bed biomass gasification expansion on industrial scales |
138 | Almond pruning | Air and steam | An increase in gasifier temperature and steam-to-biomass ratio (S/B) generated better syngas production (CO + H2) with higher hydrogen content in producer gas. Air supply in the dual fluidized-bed system had a minimal effect on biomass gasification because it did not influence the actual process | The current model does not account for tar formation, and the CFD results need to be enhanced, particularly for the cases involving CH4 and CO2. The model underestimates the prediction for CO2 by approximately 28% |
139 | Rubber wood and neem | Air | A CFD model examined how tar species (benzene, naphthalene, toluene, and phenol) formed in a downdraft gasifier through primary, secondary, and tertiary stages of tar production. Simulation results showed that CO combustion achieved the fastest reaction speed at ER 0.4 and methane formation operated at its greatest rate | Since this paper has already considered some compound presents in the tar which can be explored in further research for better understanding to reduce its formation during the gasification process |
CFD models provide the highest spatial and temporal resolution for simulating biomass gasification processes by integrating fluid flow dynamics with heat and mass transfer, as well as chemical reaction kinetics. These models serve as powerful tools for analyzing in-reactor phenomena such as flow distribution, temperature gradients, and the formation of hot or cold zones, all of which critically influence reactor efficiency and performance. Euler–Lagrangian and Euler–Euler CFD models have also been shown to be able to capture phase interactions, and predict temperature fields very accurately. For example, ZiTeng et al.140 employed a hybrid strategy where the Euler–Lagrangian method was used to model particle-level interactions at the interface of solid and fluid phases, while the Euler–Euler method captured the macroscopic behavior and interpenetration of continuous phases. Such approaches have demonstrated high accuracy in predicting temperature fields, phase behavior, and fluid–solid interactions within gasification reactors. The large number of equations describing the CFD models, particularly in the construction of multiphase systems or models with complex chemistry, is however a major bottleneck and implies that the computational requirement is high.141 Further, simplified kinetics or empirical correlations are still used by the chemical sub-models nested within most CFD studies, which limits the ability to accurately predict the formation of secondary species such as tar and soot. To address these challenges, ongoing research is exploring the development of reduced-order CFD models and hybrid modeling platforms that combine CFD with kinetic or data-driven sub-models. These emerging approaches aim to balance the trade-off between model fidelity and computational efficiency, enabling more practical and accurate simulations for reactor design, scale-up, and control. Future advances in this area could significantly enhance the predictive capability of CFD tools while maintaining manageable computational costs.
Among various data-driven techniques, the Artificial Neural Network (ANN) method is the most widely used. The ANN techniques develop predictive models by creating input–output data correlations. The methodology uses input data sets as its complete requirement thus redundancy of mathematical description is not necessary.17 ANN models are extensively used because they successfully detect complex nonlinear data connections between inputs and outputs. Among the decision-making factors the researchers employ for method selection are the application type and data access along with computational potency and desired model performance outcomes.144 This model offers the flexibility to incorporate different important process parameters such as tar content, char content, the steam-to-biomass ratio (for steam gasification), and unconverted carbon along with the other relevant variables essential for the precision of the modeling.145 There are various ANN architectures, including feed-forward back propagation neural network (FFBP), Elman-forward back propagation neural network (EFBP), cascade-forward back propagation neural network (CFBP), along with nonlinear autoregressive neural network (NARX), and layers recurrent neural network (LR). This model is optimized using algorithms such as Levenberg–Marquardt (L–M), Genetic Algorithm (GA), and Particle Swarm Optimization (PSO) methods.146
The ANN model varies in terms of architecture, reactor types, operating parameter, application focus and integration with different data sets of different gasification systems. A study was developed using ANN system which forecasts gas mixtures produced in fixed bed downdraft biomass gasifiers particularly CH4, CO and CO2 as well as H2 compositions.143 It processed elemental composition along with ash and moisture content and reduction zone temperature as inputs to reach high accuracy levels of R2 > 0.99 for CH4 and CO and R2 > 0.98 for CO2 and H2. The gas composition prediction model showed reliability through Garson's equation which evaluated the relative importance of input variables. Joel George et al.147 developed an ANN model in MATLAB to simulate the gasification process using available experimental data from the air gasification in a bubbling fluidized bed gasifier taking various biomass types. The model was trained using a multi-layer feedforward neural network with the Levenberg–Marquardt backpropagation algorithm, minimizing mean squared error (MSE) utilizing the supervised learning method. The performance analysis showed a strong agreement between predicted and the actual values, with a regression coefficient having (R) of 0.987 and the MSE of 0.71. The model effectively predicted producer gas yield based on seven key input variables, including biomass composition (C, H, O), gasification temperature (T), equivalence ratio, together with ash content (AC), and moisture content (MC). The results validate ANN modeling which is a reliable tool for the gasification process simulation. In another study, H. O. Kargbo et al.148 developed an ANN based model to optimize the operating conditions for two-stage gasification, which is aiming for high carbon conversion, together with increased hydrogen yield, and decrease carbon dioxide emissions in nitrogen and carbon dioxide rich environments. This model has aligned well with experimental data, also confirming the accuracy. Here, the optimal conditions (900 °C in stage 1, 1000 °C in stage 2, with a steam/carbon ratio of 3.8 in nitrogen and 5.7 in carbon dioxide) which resulted in the gas yields of 96.2 wt% (N2) and 97.2 wt% (CO2), with hydrogen yields of 70 mol% (N2) and 66 mol% (CO2). The carbon dioxide concentrations were minimized to 16.4 mol% (N2) and 12 mol% (CO2). Fig. 10,40 presents the basic framework for development of gasification simulation using ANN. It utilizes experimental and simulation data, various input parameters, reactor dimensions data and output parameter. ANN models simulate a given process by finding correlation between input, hidden and output layer along with reducing the mean square error compared with the experimental or simulated data available in the previous research data.
The ANN model demonstrates strong predictive capabilities for syngas composition; however, further improvements are needed to minimize prediction errors. A key limitation is the scarcity of both simulated and experimental data, which affects the model's flexibility and efficiency. Additionally, it is essential to enhance the model to account for polymeric (lignocellulosic) feedstock compositions, gasifier design parameters, and the formation of tar and char. The data driven model with physics informed neural networks (PINNs) and their derivatives known as Disentangled Representation PINNs (DR-PINNs) are recent data-driven modeling techniques that have provided solutions to predict and model gasification processes. Experimental or simulation data may be used in developing these models and physical limits (conservation of mass and energy) are imposed. Research by Ren et al.149 shows that PINNs can be used to predict syngas composition under a variety of operating scenarios with high accuracy (greater than 0.95 R2) and thermodynamic concurrency. These physics-sensitive frameworks are more interpretable and have improved generalization capabilities as opposed to black-box machine learning models. Nevertheless, there still are problems to be solved, especially in the case of handling stiff reaction systems and the extension of the models to previously impervious feedstock/reactor combinations. Moreover, stability of the models at saturation extremes and the requirement of large and good quality datasets is still a concern. Current research is tackling these shortcomings with integration of conservation laws as hard constraints and integration of PINNs with mechanistic reactor sub-models to provide additional robustness and flexibility.
Each of the four modeling approaches used to address gasification problems has its own strengths and limitations. Table 6 provides a detailed comparison, outlining the advantages, disadvantages, and key features of each modeling method.
Comparison between different modeling process133,146,150–153 | |||
---|---|---|---|
Thermodynamic equilibrium model | Kinetic model | CFD modeling | Data driven modeling |
The thermodynamic equilibrium models draw from the second law of thermodynamics by applying equilibrium constants and minimizing Gibbs free energy while optimizing entropy | The foundations of kinetic models rest upon reaction kinetics together with rate of reaction, reactor hydrodynamics and geometry | CFD modeling depends on conservation laws for mass, momentum and species together with energy within predefined boundaries | The modeling methods from data-driven approaches require experimental and simulation data for creating mathematical relationships through regression analysis, tree-based algorithms or support vector machines |
This equilibrium model is straightforward to apply and is independent of the gasifier's design | Depends on the gasifiers design | It highly depends on gasifiers designs | Since it is data driven model, so the gasifier designs are their input parameter itself |
Time invariant | Time-dependent and can predicts system changes over time | It can perform both steady and time-dependent simulates to analysis the steady and transient behavior | It can handle both static and time-series data |
It has relatively simple mathematical calculations | It is more complex due to tough differential equations need to solve for calculation | Highly computationally intensive since it requires fine mashing to capture the physical parameter accurately | It can varies, but can be computationally expensive for large datasets |
While the model provides accurate predictions of maximum yield, it tends to overestimate or underestimate the quantities of methane and char produced | Generally, more accurate for non-equilibrium conditions | This has the capability to deliver comprehensive spatial and temporal predictions | This model is completely depending on the quality and quantity of |
This is moreover limited to the equilibrium condition | It can be used for modeling non-equilibrium states and transient behavior | The CFD model is highly adaptable to different reactor designs and operating conditions | It is versatile and can be applied to various types of data and prediction tasks; however, there remains a lack of sufficient data for all types of gasifiers and gasification processes |
The main purpose of this model is to estimate syngas yield while lacking the ability to accurately model complex events including tar formation and char production, minor hydrocarbon generation and heat loss | Kinetic models use basic methods to simplify the processes involved in tar formation as well as cracking dynamics while handling complex fluid dynamics especially in fluidized bed systems. The precise modeling of char reactivity changes during the gasification process proves difficult to achieve | The computational constraints stop researchers from running direct CFD simulations that require modeling all particle and turbulence scales extending from micrometers to meters | Gasification data driven modeling faces limitations due to its requirement for good data quality and its susceptibility to overfitting as well as its inability to explain what is happening. The system faces difficulties when dealing with uncertainties while providing small information about actual physical processes |
In summary, while each modeling approach offers unique advantages, no single method sufficiently captures the full scope of physical, chemical, and operational complexity inherent in biomass gasification. A forward-looking strategy should focus on integrating these models such as coupling reduced-order kinetics with CFD insights or embedding mechanistic understanding into physics-informed machine learning models. By doing so, the field can move beyond descriptive modeling and toward predictive, scalable tools that support reactor optimization, techno-economic evaluation, and commercial deployment.
CO + 3H2 ![]() | (R.15) |
CO2 + 4H2 ![]() | (R.16) |
The methanation reactors usually face three major challenges like the presence of catalyst poisons (especially the sulfur compounds), carbon deposition, and the highly exothermic nature of the reactions.165 To ensure a catalyst lifespan of at least one year, which is considered economically viable, the sulfur content present in the feed gas must be kept at below 1 ppm.166 Secondly, the carbon deposition is particularly problematic at temperatures above 500 °C. Since CO absorbs strongly on nickel-based catalysts, which can decompose into carbon atoms. If it is not rapidly hydrogenated, this carbon can lead to the formation of polymeric carbon with carbon nanofibers, nickel carbides or, all of these can block active catalytic sites.167 Additionally, the exothermic nature of reaction causes significant temperature increases within the reactor, which potentially can create a hot spot between 550 °C and 750 °C, which can result in the catalyst sintering and loss of activity.
Moreover, the modeling of methanation reactors relies also on fixed-bed, fluidized-bed systems and specialized designs including fixed-bed tube-bundle and structured fixed-bed reactors. The models integrate multiple elements which include time resolution alongside reactor dimensions, phase representation and temperature behavior together with kinetic approach and software simulation tools. The researcher needs to boost the efficiency of methanation combined with improved catalyst performance and sustainable operations while generating pure CH4 during low-temperature extended operations. Currently operating methanation systems exist commercially yet researchers must conduct additional studies about catalyst mechanisms and feed gas composition variations. The research emphasis stands on catalyst enhancement while working to develop better reactor models with optimal temperature control systems. There is necessity to focus on improving three main performance areas that include operational flexibility and dynamic performance with cost efficiency.
Gasification technology is basically a reliable and green source to produce hydrogen since we can utilize various biowaste to convert into rich hydrogen syngas. The recent technologies for producing hydrogen from the biomass include pyrolysis with gasification and various methods for converting biomass into liquid fuels such as hydrolysis, along with liquefaction and supercritical extraction. Sometimes these processes are followed by reforming to enhance the generation of hydrogen yield.173 When air is used in the gasification system, it can produce a gas mixture that contains approximately 20% hydrogen, 10% carbon dioxide, 5% methane, along with 20% carbon monoxide and 45% nitrogen. This gas stream can be modified further through a reaction with steam to convert the carbon monoxide into additional hydrogen utilizing the water–gas shift reaction.174 Whereas supercritical water gasification is also an efficient thermochemical process which enables moist biomass feeding directly to the gasifier along with pressure-induced hydrogen storage cost reductions.175 Another method to produce hydrogen via biomass gasification with calcium oxide (CaO) serves as a potential viable method for hydrogen generation. This eco-friendly biomass gasification approach enables large-scale production of hydrogen while utilizing widely available and affordable CaO catalyst to generate hydrogen-rich gas products.176 While the solar based gasification technology is reliable as it provides green source of power to the gasifier. This can also improve the effectiveness of feedstock and overall energy by 30% and 40% respectively.177 The chemical reactions below illustrate different pathways for syngas reformation produced from biomass gasification. This syngas, composed mainly of CO and H2, is directed into a turbine, where it generates power under high-temperature and high-pressure conditions. After exiting the turbine, the syngas passes through a heat exchanger, where it is cooled to a suitable temperature for water heating, enabling steam generation for the electrolysis process. Subsequently, the syngas enters a multi-stage water–gas shift reactor (MWGSR), where steam is used to convert CO into CO2 while producing additional hydrogen, as represented by reactions (R.17) to (R.20).28 Additionally, a portion of the syngas can be directly combusted in a Brayton cycle to produce power.178
Biomass → char + C6H6 + CO + N2 + CH4 + H2O + H2S | (R.17) |
Char → C + O2 + N2 + H2 + S + ash | (R.18) |
C + H2O → CO + H2 | (R.19) |
CO + H2O → CO2 + H2 | (R.20) |
Owing to high carbon dioxide production, hydrogen production by steam reformation is classified as gray hydrogen. Hydrogen production coming from sources like natural gas, biogas or syngas is classified as blue hydrogen. In case of blue hydrogen, the CO2 emissions can be brought down through carbon capture methods and subsequent reuse practices while grey hydrogen plans lead to atmospheric carbon release. Biomass gasification produces environmental benefits through lower greenhouse gas emissions that range between 405–896.61 g CO2 per kg H2 while wind-powered electrolysis produces 600–970 g CO2 per kg H2 emissions.179,180 The worldwide hydrogen manufacturing amounts to 75 million tons per year split into blue hydrogen produced through natural gas that uses 205 billion cubic meters of gas from natural sources while making up 6 percent of all gas use on a global scale. A total of 23% represents grey hydrogen fuel which is derived from coal to generate 107 million tons corresponding to 2% of worldwide coal consumption.181 Moreover, the efficiency level for hydrogen gas production can be improved by integrating biomass torrefaction with densification and gasification operations.
N2(g) + 3H2(g) → 2NH3(g)ΔH = −92k, Haber–Bosch reaction | (R.21) |
Fig. 12 describes the two-process integration methods to produce green ammonia utilizing renewable sources like syngas gas produced from biomass and renewable power sources inside the electrolysis. Green ammonia functions as a storage solution and transport method for hydrogen to resolve hydrogen storage and transportation issues. Whereas the cracking process transforms stored ammonia back to hydrogen along with nitrogen by applying heat (750–850 °C) and a proper catalyst which is again a costly process since it requires high heat energy. Through biomass gasification processes agricultural waste together with crop residues turn into syngas by transforming into H2, CO2, H2O, CH4, CO, N2 and air mixture. The synthesized syngas provides a suitable raw material for ammonia production which substitutes fossil fuels while enabling environmentally friendly ammonia manufacturing.186 Weng et al.187 have simulated hybrid biomass conversion process into ammonia by using chemical looping with solar and wind power system. This system consists of a biomass gasification system, chemical looping air separation (CLAS), water electrolysis with chemical looping ammonia production (CLAP) and power generation sources. Parametric optimization and feasibility assessments were carried out using Aspen Plus simulations. The cascading utilization of the biomass made the simultaneous production of N2, H2 and NH3. The standard operating conditions with a biomass feed rate of 1 kg s−1, made system to achieve an ammonia selectivity of 79.36%, with an ammonia concentration of 65.65 vol% and a production rate of 34.1 kmol h−1. These outcomes demonstrate the viability of the HBCAS approach and offer valuable insights for the practical application. Another study, Nejat Tukenmez et al. modeled a multigeneration plant based on solar and biomass power generation. It has utilized a gasifier, hydrogen compressor, cooling unit, parabolic dish collector, Rankine cycle, ammonia storage tank, ORC cycle, hot water production unit, along with PEM electrolyzer and ammonia reactor unit. This integrated plant's total electrical energy output has been determined to be 20125 kW. Whereas its energy efficiency with exergy efficiency is evaluated at 58.76% and 55.64%, respectively. And the production rates of hydrogen and ammonia are calculated to be approximately 0.0855 kg s−1 and 0.3336 kg s−1, respectively.
The production of green ammonia utilizes alternative sources of renewable energy consisting of solar power and wind energy as well as hydroelectric power to operate electrochemical reactors integrated with biomass gasifier which has potential to decrease carbon pollution. Where the prices for electrolytic ammonia production stands at $680–900 per ton but experts predict, it will drop to $400 per ton by 2030.188 The widespread implementation of ammonia as a commercial fuel faces difficulties because of its low energy content alongside high ignition requirements and NOx emissions when burned. The main obstacle in this method occurs when researchers seek an efficient catalyst to perform ammonia synthesis under low temperature and pressure conditions. A suitable catalyst plays a vital role in overcoming the high energy requirements of nitrogen reactions while improving the process viability.189,190 The stationary power sector currently accepts ammonia as a renewable fuel with support from extended purchase agreements. The advancement of technology is anticipated to lead to better electrochemical processes while reducing costs which will make its implementation more possible.
- Carbon efficiency: PBtX systems can achieve carbon efficiencies exceeding 90%, which is significantly higher compared to 25–40 percent reported for conventional biomass-to-X systems.191
- Product yields compared to dry biomass feed: recent studies have reported that such systems can produce between 0.31 and 0.79 kg of syngas, 0.70 to 1.28 kg of methanol, and 0.20 to 0.57 kg of Fischer–Tropsch (FT) liquids per kilogram of biomass feedstock.192
- Efficiency of electrolyzer: electrolyzer efficiency is another critical determinant of PBtX system performance. Low-temperature electrolyzers such as alkaline electrolysis (AEL) and proton exchange membrane (PEM) systems typically operate at efficiencies of 50–58%, while solid oxide electrolysis cells (SOECs) can theoretically achieve up to 80% efficiency under optimal conditions.193
- Specific capital costs: integrated PBtX systems demonstrate approximately 30% lower capital expenditure than standalone PtX plants, with reported costs around €3580 per kWlhv, this is large due to yield improvements achieved by system level integration.191,194
The economic feasibility of PBtX systems, even with the substantial efficiency metrics, is very dependent on electrolyzer capital expenditures and the costs of renewable electricity. Research highlights that firstly to be competitive electrolyser capacity factors need to be maintained above 82–94%.191 This requirement is difficult to meet when relying solely on intermittent renewable energy sources, which necessitate backup power solutions or energy storage infrastructure, both of which introduce additional capital and operational expenditures. Secondly, economic modeling shows that PBtX systems based on surplus electricity alone cannot be made to work unless they are kept in non-stop operations. Biomass gasification offers stabilization, allowing a more predictable utilization of capacity. However, ensuring a reliable biomass supply chain and managing the complex interface between electricity fluctuations and gasification operations remain persistent technical barriers.195 The third is efficiency losses & heat management, at theoretical efficiencies of nearly 80%, SOEC-enhanced systems hold promise, but, in many instances, device efficiencies are much lower as presented by the auxiliary loads and system losses. As an example, methanol PBtX chains may have declined exergy efficiencies of ∼70 to ∼58% under real conditions.193 Hence, a combination of heat recuperation and process maximizing is a priority. Finally, is Technology Maturity & Scale-Up Risk. Most of PBtX ideas remain on simulation or pilot phase. Demonstrations on the TRL 7–8 scale are necessary to confirm integration of electrolysis into a gasification system, dynamically testing at variable electricity, syngas conditioning and removal of tar systems.196 Otherwise, commercial investors will be hesitant about investing in it when there are doubts about the scalability of it. The further possible advances to improve the PtX system discussed in Table 7 in terms of CAPEX, efficiency, integration, scaling up infrastructure and favorable policies.
Challenge | Required advances |
---|---|
CAPEX reduction | Scaling up the electrolyzer/SOEC capacity; homogeneous biomass gasification system; supply chain optimization to lower logistics and shared infrastructure expenses |
Efficiency under flexibility | Electrolyzers/gasifying reactors that can be partially loaded and cycled without any performance drawback; better thermal integration |
System integration | Achieving optimal integration of electrolysis, biomass gasification, and scaled-up renewable electricity infrastructure; completing full process integration including water–gas shift (WGS) and synthesis; and developing advanced syngas purification techniques adaptable to dynamic operating conditions |
Demonstration scale | Pilot plants can be established to demonstrate Power-and-Biomass-to-X (PBtX) pathways, and to validate Technology Readiness Levels (TRL) 7–8 through comprehensive techno-economic analysis and life cycle assessment (LCA) |
Policy & incentives | Increases in carbon prices, renewable hydrogen credits and subsidies to high-capital PtX infrastructure to fill in cost gaps |
![]() | ||
Fig. 13 Percentage of renewable energy in energy mix in 2050 by four major stakeholders.198 |
A study evaluated the thermodynamic and economic feasibility of three synthetic natural gas (SNG) production systems based on biomass gasification integrated with syngas methanation. The three case scenarios included case1 as dual fluidized bed (DFB) gasification with CO2 capture (DFB + M + CCS), case 2 as DFB with renewable hydrogen (DFB + EL + M) and case 3 as direct biomass gasification with renewable hydrogen (CFB + EL + M). Case 3 achieved the best performance by recovering over 98% carbon content while maintaining efficient cold gas performance at 77.10% and exceeding the results of Case 2 at 70.92% and Case 1 at 63.27%. A better heat integration approach enabled the system to achieve 14.59% enhanced efficiency performance. To become financially competitive the SNG break-even price for Case 2 should remain between 58–98 € per MWhSNG while renewable electricity stays below 57.9 € per MWh. The operational model for Case 3 could thrive when electricity prices remained below 70.5 € per MWh.200
Broadly hydrogen is divided into different varieties namely blue, gray, brown, black and green depending on how it is produced, the energy source and environmental effect it generates. Renewable hydrogen is expected to achieve $225.55 billion in the market by 2030 while maintaining a 6.4% compound annual growth rate. By 2026 the U.S. government wants to produce hydrogen at $2 per kilogram through efficient low-carbon technology solutions which they expect to achieve $1 per kilogram production by 2031. The currently employed thermochemical along with electrochemical, biological and photolytic processes do not meet the target cost rates.201 A study performed on two process models to generate hydrogen and syngas from both coal and natural gas.193 The first case employs literature-validated entrained flow gasification facilities though Case 2 implements a reforming system which improves hydrogen generation and minimizes carbon output. The Case 2 process achieves an 88% enhanced HCR value of 1.20 which surpasses baseline values by 1.20. Moreover, the two-process model offers 55% more hydrogen production along with an 18.5% enhancement in operational efficiency. Case 2 reduces carbon emissions by 69.6% for each unit of produced hydrogen. The investment for producing each ton of hydrogen and the final hydrogen selling price in Case 2 fall at levels which are 28.9% below Case 1 rates. For each metric ton of fuel, the total production cost amounts to €1892.092 in Case 1 and €1344.984 in Case 2. The calculation of total production cost per ton indicates that the TPC amount in Case 2 shows a lower value than the TPC amount in Case 1. The study demonstrates that the minimum selling price of hydrogen operates at a lower competitive level in Case 2 versus Case 1. The equation is used to calculate the fixed cost like equipment cost of process design and the total investment cost per ton of H2 as mentioned in eqn (16) and (17).
![]() | (16) |
![]() | (17) |
Biofuel sectors face technological challenges along with economic hurdles in generating hydrogen, especially in mini decentralized facilities. Research into a 100 kWth system evaluated production costs by analyzing capital investment expenses and running costs together with efficiency parameters. The research demonstrates that efficiency boosts can decrease total expenses but must be strengthened through additional methods. The Portable Purification Unit (PPS) cost stands as the main cost factor because it requires significant reduction. The specific production cost will fall between 9.5 to 12.75 € per kg when operational costs decrease by 50% and the steam-to-biomass ratio goes from 1 to 1.5.202 Therefore, the sensitive part of techno-economic analysis for hydrogen production varies by process. Electrolysis studies focus on capital, operating, maintenance costs, and net present value (NPV), while reforming considers production costs, including reactors, membranes, and labor. Gasification mainly evaluates NPV.
Ammonia (NH3) is synthesized from nitrogen and hydrogen using the Haber–Bosch process, with hydrogen production being the primary economic challenge. A study evaluated the technological feasibility together with economic assessment for producing ammonia through biomass gasification in a pulp and paper facility. Within the integrated system the overall energy efficiency increased by 10%-units above a traditional ammonia production plant. The economic viability needed an increased selling price range of 509–774 € per tons NH3 to reach investment return rates between 10–20%. Investment costs representing 45% were allocated to the synthesis loop. The evaluated production capacity at 228000 tons per year proved unprofitable for current market prices. Plant expansion and cost reduction or increased prices from fossil-based alternatives establish the economic basis for financial feasibility.203 A technological and financial analysis evaluates the two methods of producing green ammonia against traditional methane-to-ammonia processes alongside heat integration and improved steam cycle management. The 50k ton per year reference production shows different patterns between operational efficiency and cost expenditures. Power-to-ammonia demonstrates the best efficiency at 74% while methane-to-ammonia operates at 61% efficiency, yet power-to-ammonia achieves only 44% efficiency. The biomass-based production methods require high costs at $450 per ton and need more than six years to break even whereas methane-based products cost $400 per ton and reach their payback in five years. The power-to-ammonia process exists at present only as an uncompetitive method but advanced solid-oxide development together with expanded renewable energy usage may lead to its future viability.204
The hydrogen production technologies show that switching from natural gas to biomass as a source reduces GHG emissions by approximately 75%.207 The analysis of popular gasification process demonstrates that the transformation method presents an alternative option to conventional natural gas reforming when evaluated at various points in its life cycle. This method reduces GHG emissions by 0.4 kg CO2e per kg H2 more efficiently than natural gas reforming which results in 10.6 kg CO2e per kg H2 emissions and lowers fossil fuel usage.208 Pyrolysis is also one of the most environmentally friendly methods for hydrogen production. The process produces hydrogen-rich syngas by thermally decomposing biomass in an oxygen-deficient environment.209 The LCA of the process revealed that it generated decreased greenhouse gases, water usage and energy requirements than alternative hydrogen production approaches. The production of hydrogen through pyrolysis costs almost 2$ per kg of H2 production and has 40% of recovery efficiency. Thus, the gasification requires both higher temperatures along with oxygen utilization and produces larger environmental effects than the pyrolysis does. In contrast, dark fermentation generates hydrogen and organic acids from biomass under anaerobic conditions but incurs higher environmental costs compared to the pyrolysis process. Moreover, the steam reforming of biomass demonstrates high-energy requirements and elevated emissions during its processing as a method for hydrogen production. The LCA research demonstrates that using biomass to make hydrogen produces better environmental results than both natural gas reforming and water electrolysis processes. Mass production efficiency from pyrolysis serves as an affordable solution having better performance than gasification, dark fermentation, and steam reforming at reducing energy usage and water demand.210
Studies have also examined the various hydrogen production methods based on environmental impacts, including carbon dioxide equivalent emissions, acidification, eco-toxicity, along with the human toxicity, carcinogens, and abiotic depletion.211 Investigations revealed that the global warming potential of steam methane reforming and coal gasification was 3.03 and 3.85 kg CO2-eq per kg of ammonia produced, respectively. In contrast, biomass gasification had the lowest global warming potential at 0.378 kg CO2-eq per kg of ammonia. Additionally, abiotic depletion was highest for steam methane reforming, measured at 0.0264 kg Sb-eq per kg of ammonia produced. Fig. 14 presents the greenhouse gas emissions associated with various ammonia production methods. It indicates that ammonia produced from methane results in the highest CO2 emissions, while the gasification-based process exhibits the lowest emissions.211
Furthermore, the analysis demonstrates that municipal waste incineration together with hydropower techniques produces ammonia with lower environmental impacts when compared against alternative production methods. The quantified greenhouse gas (GHG) emission levels for one kilogram of produced ammonia amount to 0.34 kg CO2-eq using municipal waste and 0.38 kg CO2-eq using hydropower and 0.84 kg CO2-eq from nuclear facilities as well as 0.85 kg CO2-eq from biomass generation. The study also performed an energy and exergy analysis to determine sustainability index scores which represented opportunities for improvement. Four types of systems examined for ammonia production efficiency resulted in energy efficiencies of 42.7% for hydropower while nuclear reached 23.8% and biomass achieved 15.4% and the lowest 11.7% was achieved by municipal waste-based systems. The calculated exergy efficiencies for energy production reached 46.4% for hydropower and 20.4% for nuclear and 15.5% for biomass systems while municipal waste-based methods had an exergy efficiency of 10.3%.212,213 Another study investigated the process integration of the gasification system at olive oil mills with heat and power production through CHP with biochar manufacturing that uses moist olive pomace as fuel source. The system generates electricity with 13.5% efficiency and 32% CHP efficiency which produces 0.88 kW h renewable electricity per kg of olive oil. LCA analysis reveals an 8.25% decrease in environmental effects while achieving a 21% reduction of climate change emissions that fall from 2.21 to 1.74 kg CO2eq per kilogram of olive oil. The proposed gasification plant provides responsible methods to manage olive pomace while recovering sustainable energy.214 Therefore, biomass gasification technologies have the potential to reduce the environmental impact by reducing the emission of carbon and utilizing it for the formation of methane, ammonia and hydrogen. Sustainability assessment including economic viability, such as different biomass gasification technology with its energy efficiency, economic metric emission and economic feasibility considering its cost and technology readiness levels (TRL) scale discussed in Table 8.
Ref. no. | Technology | Energy efficiency | Economic metric | GHG emissions/reduction | CAPEX/OPEX & payback | TRL/scale |
---|---|---|---|---|---|---|
215 | Biomass fluidized-bed (FB) | 40–50% (LHV) | LCOE (levelized cost of electricity) ≈ 0.067 USD per kWh in China | Near-carbon neutral | Moderate CAPEX; low OPEX | TRL (technology readiness levels) 9, commercial |
216 | Entrained-flow biomass → H2 | ∼56% (LHV) | Min H2 price ≈ 150 € per t | Negative lifecycle GHG possible | High CAPEX; needs ≥150 € per t H2 to break even | Pilot/demo |
217 | Biomass chemical-looping gasification → FtL | 52–53% overall biomass to liquid conversion efficiency chain | BESP (Breakeven selling price) ≈ € 781–816 per m3; CO2 avoidance payback | GHG reduction 79% (no CCS); up to 264% with CCS | BESP €781–816; competitive with heat valorisation & CO2 credits | Demo-scale BtL |
218 | Medical-waste plasma + SOFC + S + CO2 cycle + desalination | 41.7% net power; 65% energy utilization rate | Low investment −62![]() |
Waste CO2 fixed ∼59.8k t per year | Payback 4.4 year; NPV $118 M during 20 years of life span | Conceptual/system-level |
219 | Plasma gasification (general waste-to-energy) | Thermal efficiency 60–80%, net electrical efficiency 18.6–25% | Operating capacity 1000 t per day with net output 33 MWH | Varying; overall lower pollutants than incineration | CAPEX 250–450 M USD for 2000–3000 t per d; cost per t drops with scale | Demo-to-commercial |
220 | Solar-driven biomass gasification hybrid (PV + solar thermal) | 65% total energy conversion | Fuel cost = 354 $ per ton methanol (market range) | −0.56 kg CO2eq per kg methanol (net reduction) | Fuel cost within market; integrated PV reduces system expenses | Model-to-pilot scale |
221 | SCWG of sewage sludge for H2 + power | System thermal efficiency ≈ 63% | Fixed capital cost = 8 millions CNY (sewage sludge treatment capacity 20 t per h), LCOE-$0.696 per kWh | Environmentally favorable; subsidies boost ROI | H2 and power; ROI ≈ 24.9%, PP ≈ 3.8 year; H2 cost ≈ 17.07 CNY per kg | Pilot-to-demonstration |
222 | Multistage biomass FLETG (fixed-bed, low-emission, two-stage gasifier) type gasifier | ∼81% gasification efficiency; HHV ∼6.4 MJ Nm−3 | — | Complex reactor, pilot to demo scale | Complex; reactor costs higher; pilot scale exists | 100 kW to 6 MW pilot/demo |
Moreover, there is much potential to improve the gasifier technology to get better cold gas efficiency and carbon conversion efficiency. The updraft gasifier faces the main challenge of producing gas with high tar levels that reduces its heating capacity. Researchers need to explore new methods to remove unburnt carbon in entrained-flow gasification since the approach fails to eliminate the unburnt carbon under all experimental settings. The plasma gasification method also requires enhanced modeling techniques to enable reliable operation and optimize overall performance, including maximizing plasma heat utilization and achieving uniform heating. Also, the commercial application of supercritical water gasification faces difficulties at industrial scale because of varying biomass content and operating parameters. Research into process improvement remains crucial because biomass variations lead to different responses during different operating temperatures and pressures as well as reaction times and feedstock concentrations and catalyst types. The Concentrated Solar Thermochemical Gasification of Biomass (CSTGB) system increases biomass utilization and power efficiency through producer gas storage by 30% and 40%. Whereas technology faces economic obstacles despite these facts so stakeholders must implement incentive-based policies to overcome this challenge. Practical implementation of the CSTGB process requires additional studies regarding pilot-scale economic models and solar collector materials as well as heat transfer fluid technologies.225 Therefore, during the conversion process biomass-based systems generate energy outputs which fall 20–70% below what conventional natural gas steam reforming produces.226 Furthermore, syngas derived from biomass gasification exhibits low H2 to CO ratios that cause energy usage in upgrading systems to increase substantially because it requires energy-intensive units like water–gas shift reactors along with CO2 removal systems using amines and PSA. Composition of syngas depends on both gasification temperature and steam-to-biomass ratio, but gasifier technology stands as the main determining element. And there is requirement of complete shifting towards renewable energy and building large infrastructure for green energy production because the energy cost for production of green fuel like H2, CH4 and NH3 is very high from our conventional resources. There are possibilities to reduce the total capital cost by utilizing the biomass preprocessing methodology integrating with green fuel production system. This can reduce the heterogeneous biomass handling cost and various capital costs.
1. Modern gasification technologies have shown efficient performance, with solar gasification standing out for its significant potential. By utilizing renewable energy sources, it can achieve temperatures exceeding 1300 K, with energy efficiencies ranging from 70.6% to 72.7%. Despite challenges related to financial feasibility, solar gasification holds considerable value with the potential to minimize possible environmental hazards.
2. Modeling methods are fundamental to improve gasification operations and estimate syngas production from biomass. Equilibrium model delivers easy computational operations that predict thermodynamically maximum gas yields through its simple calculations, but kinetic models combine enhanced precision with time-responsive solutions which simplify tar formation processes and fluid movement operations. Among modeling techniques CFD proves most flexible because it can determine flow patterns as well as heat distribution and critical areas. Additional development of these gasification models will be essential for boosting efficiency together with accuracy and general usability.
3. Power production from green fuels, such as hydrogen, methane, and ammonia, offers significant potential for reducing carbon emissions across various industries. The methanation process can help to improve the syngas quality, with potential to reduce CO and CO2 composition by 50% and 40% respectively. Whereas, ammonia has proven to be an important carbon free hydrogen carrier, and it is expected to become economically viable by 2030, with production costs projected to drop to $400 per ton, down from the current range of $680–900 per ton.
4. The overall cost of the gasification process is highly sensitive to electricity prices, particularly in electrochemical syngas generation, which demands substantial green sources of electrical energy. However, integrating direct biomass gasification with renewable hydrogen can significantly enhance efficiency—recovering up to 98% of the carbon content while maintaining a cold gas efficiency of 77.10%. This approach allows electricity costs to remain below 70.5 € per MWh. Moreover, the use of green hydrogen not only contributes to a more sustainable process but also helps reduce the production costs of methanation and ammonia.
5. Transitioning from natural gas to biomass as a feedstock for hydrogen production can significantly reduce greenhouse gas emissions by approximately 75%. Among the evaluated technologies, biomass gasification demonstrated the lowest global warming potential, with an emission value of just 0.378 kg CO2-equivalent per kilogram of ammonia produced. Integrating renewable energy and biomass gasification to produce green fuel in a multigeneration plant offers a promising pathway toward achieving negative carbon emissions.
ANN | Artificial neural network |
CFD | Computational fluid dynamics |
CGE | Cold gas efficiency |
DNS | Direct numerical simulation |
GHG | Greenhouse gas |
EDGAR | Emissions Database for Global Atmospheric Research |
ER | Equivalent ratio |
GHG | Greenhouse gas |
IEA | International Energy Agency |
IRENA | International Renewable Energy Agency |
LCA | Life cycle assessment |
LES | Large eddy simulation |
LHV | Low heating value |
MSW | Municipal solid waste |
MHGCG | Multistage heating and gradient chain gasifier |
NPV | Net present value |
RANS | Reynolds–Averaged Navier–Stokes |
PBtX | Power and Biomass-to-X |
SCWG | Supercritical water gasification |
SSF | Simultaneous saccharification and fermentation |
SMR | Steam methane reforming |
SOE | Solid oxide electrolysis |
SNG | Synthetic natural gas |
SOEC | Solid oxide electrolysis cells |
SOFC | Solid oxide fuel cell |
TEqM | Thermodynamic equilibrium models |
TRL | Technology readiness level |
α | Reaction order |
αg | Gas volume fraction |
β | Temperature exponent [1/K] |
g | Gas density [kg m−3] |
μi | Chemical potential of species i |
μg | Dynamic viscosity [Pa s] |
τg | The sum of viscous stress and Reynolds stress [s] |
Γeff | The sum of molecular and turbulent heat diffusion coefficients [m2 s−1] |
λi | Lagrange multipliers |
λg | Gas thermal conductivity [J (m−1 s−1 K−1)] |
This journal is © The Royal Society of Chemistry 2025 |