Next Article in Journal
Modelling Energy Consumption and Energy-Saving in High-Quality Olive Oil Decanter Centrifuge: Numerical Study and Experimental Validation
Next Article in Special Issue
Cost and Profitability Analysis of a Prospective Pennycress to Sustainable Aviation Fuel Supply Chain in Southern USA
Previous Article in Journal
SmartLVGrid Platform—Convergence of Legacy Low-Voltage Circuits toward the Smart Grid Paradigm
Previous Article in Special Issue
Experimental Investigation of Performance, Emission and Combustion Characteristics of a Common-Rail Diesel Engine Fuelled with Bioethanol as a Fuel Additive in Coconut Oil Biodiesel Blends
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High Quality Syngas Production with Supercritical Biomass Gasification Integrated with a Water–Gas Shift Reactor

1
School of Mechanical Engineering, University of Adelaide, Adelaide, Australia
2
Division of Computational Physics, Institute for Computational Science, Ton Duc Thang University, Ho Chi Minh City, Vietnam
3
Faculty of Electrical and Electronics Engineering, Ton Duc Thang University, Ho Chi Minh City, Vietnam
4
Department of Mechanical Engineering, Lamar University, Beaumont, TX 77705, USA
5
Sustainable Management of Natural Resources and Environment Research Group, Faculty of Environment and Labour Safety, Ton Duc Thang University, Ho Chi Minh City, Vietnam
*
Authors to whom correspondence should be addressed.
Energies 2019, 12(13), 2591; https://doi.org/10.3390/en12132591
Submission received: 30 May 2019 / Revised: 27 June 2019 / Accepted: 1 July 2019 / Published: 5 July 2019
(This article belongs to the Special Issue Research Advances in Liquid Biofuels)

Abstract

:
A thermodynamic assessment is conducted for a new configuration of a supercritical water gasification plant with a water–gas shift reactor. The proposed configuration offers the potential for the production of syngas at different H2:CO ratios for various applications such as the Fischer–Tropsch process or fuel cells, and it is a path for addressing the common challenges associated with conventional gasification plants such as nitrogen dilution and ash separation. The proposed concept consists of two reactors, R1 and R2, where the carbon containing fuel is gasified (in reactor R1) and in reactor R2, the quality of the syngas (H2:CO ratio) is substantially improved. Reactor R1 is a supercritical water gasifier and reactor R2 is a water–gas shift reactor. The proposed concept was modelled using the Gibbs minimization method with HSC chemistry software. Our results show that the supercritical water to fuel ratio (SCW/C) is a key parameter for determining the quality of syngas (molar ratio of H2:CO) and the carbon conversion reaches 100%, when the SWC/C ratio ranges between two and 2.5 at 500–1000 °C.

Graphical Abstract

1. Introduction

The energy crisis and environmental pollution due to the combustion of carbon-containing fuels have resulted in more research being conducted on renewable energy and clean energy resources as potential alternatives to fossil fuels [1,2,3]. Synthetic gas (syngas) is a mixture of hydrogen and carbon monoxide, which is a promising replacement for fossil fuels. It is a cheap, clean burning fuel, easy-to-transport gas, and has a wide range of applications such as some types of fuel cell systems [4,5] and the Fischer–Tropsch process [4]. Therefore, it has received ever-increasing special attention throughout the past decades. Gasification is the common pathway to produce syngas from carbonaceous fuels. In conventional air-blown gasification systems, air is used for the partial oxidation of fuel, which not only adds impurities (e.g., sulphur dioxide) to the final gaseous products but supplies nitrogen to the syngas, which reduces the quality of the syngas product. One plausible solution for this challenge is to use oxy-fuel gasifiers to avoid the appearance of impurities and nitrogen at the outlet, which in turn adds to the cost and complexity of the process. Therefore, there is a need to seek alternatives that produce high quality syngas, while addressing the aforementioned challenges [5].
Chemical looping gasification is an emerging technology for the production of syngas using a solid oxygen carrier (OC). This technology addresses the nitrogen dilution but also has the potential to reduce and/store greenhouse gases (GHG) such as CO2. In this concept, two interconnected reactors, the gasifier and the air reactor, are employed. In the gasifier, metal oxides are reduced, and fuel is partially oxidized [6]. Syngas is the main product of the gasifier. Then, the reduced metal oxides are transported to the air reactor where particles of oxygen are recovered [7]. Notwithstanding the advantages of the solid oxygen carrier particles, there are some challenges associated with the use of the solid oxygen carriers in chemical looping systems. These include agglomeration and sintering [6,8,9,10,11,12,13,14,15,16], and also the need to separate the OC particles from any carry-over particles from the gasifier, as well as manage the deposition of carbon and ash on the OC particles [17]. These challenges significantly reduce the effectiveness of the oxygen carrier particles to transport oxygen between the reactors [16,18,19,20,21,22] and hence decreases the efficiency of the process [6]. However, this emerging concept is in the early stage of development and needs further investigation to understand its shortcomings. For example, a material constraint due to the sintering, breakage, and corrosion of metals, is one of the challenges associated with the use of molten metal in the chemical looping process.
One promising method to gasify a carbonaceous fuel is supercritical water (SCW) gasification. This concept offers a wide range of benefits over the other concepts such as:
  • Supercritical water has zero surface tension and most of the carbon-containing fuels are soluble in it, and therefore diffusion and penetration of water in the carbon with insignificant mass transfer resistance is plausible [23];
  • The SCW gasification process is flexible with respect to the type of the carbonaceous fuel. For example, different types of biomass, coal, or even municipal waste with various contents of moisture and impurities are used as fuel sources [24];
  • The required operating temperature for the gasification lies between 400 °C and 1000 °C depending on the type of the fuel and quality of the syngas;
  • The produced carbon dioxide easily separates from H2 using pressurized water;
  • Some physical properties of water, such as density, ion product, dielectric constant, viscosity, diffusivity, and solubility, change near or at its thermodynamic critical point (T = 374 °C and P = 22.1 MPa). At the critical point, water behaves similar to a dense gas with a consequent removal of any interphase mass transport processes. Organic compounds have high solubilities and complete miscibility with supercritical water [25];
  • The process is high pressure, which reduces the costs related to the storage of the gaseous products such as post-compression operation.
These advantages provide plausible conditions for better gasification of carbonaceous fuel, particularly biomass in supercritical water. For example, thermodynamic assessment and system modelling for the gasification of biomass with supercritical water were conducted by Withag et al. [26]. The selected feedstock was a wet biomass comprised of 70% water by weight. It was shown that the gasification is feasible without any further drying process. Thermochemical equilibrium analysis was used for the modelling of the process and it was found that the composition of the product gases could be tailored to the desired product composition by changing the process parameters such as the reactor temperature, pressure, and the concentration of organic material in the feed. However, the pressure of the reactor was 100–300 bar, which was technically challenging and expensive. In addition, the proposed system produced 38% CO2, which was one main challenge of the supercritical gasification at higher temperatures.
In another work, Guan et al. [27] investigated the gasification of algae nanochloropsis in supercritical water and showed that with an increase in the temperature of the gasifier, more carbon dioxide is produced. By reducing the operating temperature of the gasifier, the mole fraction of CO2 slightly decreased, however, the quality of the produced synthetic gas (H2:CO molar ratio) was low. A similar trend was reported in the gasification of dry starch as biomass conducted by Yakaboylu et al. [28]. Thus, to achieve high quality syngas, there is a need to separate the CO2 from other gaseous products. In a study conducted by Guo et al. [29], they reported the controversial result that operating temperature has a strong influence on the gasification of biomass with supercritical water such that the gasification efficiency and H2 production at higher temperatures inverses the quantity of the CO2 product [29]. Thus, further investigation on the role of temperature on the supercritical gasification of biomass and mole fraction of gaseous products is required.
The diversity of components in gaseous products obtained by the gasification pathway has also been a popular subject of research, because carbon-containing fuels such as biomass have different compositions of cellulose, glucose, glycerol, lignin, and phenolic which may result in a substantial change in the gasification reactions and consequently changes the configuration of the reactor [29,30,31]. For example, Guo et al. [32] performed an experimental investigation on the catalytic and noncatalytic supercritical gasification of glycerol in a tubular quartz reactor for hydrogen production. They showed that by using a catalyst, the main gaseous products are hydrogen with the mole fraction of 59%, followed by CO2 with the mole fraction of 29.9%, CH4 with the mole fraction of ~6.5%, and CO with the mole fraction of ~4.5%. Noticeably, in the absence of the nickel catalyst, the mole fraction of hydrogen decreased to 50%, the mole fraction of CO2 decreased to 24.93%, and the mole fraction of CO decreased to 21.13%. Therefore, depending on the type of feedstock, a specific type of the reactor or configuration of the process is required [33,34,35].
For example, one potential configuration for supercritical gasification is a fluidized bed system. In a study conducted by Lu et al. [36], the hydrodynamic behavior of a supercritical gasifier was evaluated at temperatures ranging from 360 °C to 420 °C and pressure ranging from 23 MPa to 27 MPa. They identified that a double symmetric feeding pipe with an angle of 45° provided uniform solid distribution and a long residence time, which potentially also represented better chemical efficiency. However, it was identified that depending on the composition of the feedstock, the configuration of the system might need a major modification in order to produce high-quality syngas [33]. Hence, in our research, the thermodynamic potential for a new configuration that would produce a high quality syngas with supercritical water is investigated for graphite (pure carbon) as a surrogate for any carbonaceous fuel and biomass. The presence of the water–gas shift reactor removes the barrier of dependence of the configuration of supercritical gasification on the composition of feedstock. Thus, the proposed gasification plant is combined with a water–gas shift (WGS) reactor to control the H2:CO ratio. The influence of different operating conditions, including the ratio of supercritical water to feedstock, temperature, and the pressure of the reactors on the gaseous products is investigated. The chemical performance of the proposed concept is investigated for three different biomass feedstocks.

2. Methodology

Figure 1 presents a schematic diagram of the operating conditions considered for this research. To apply the thermochemical equilibrium and sensitivity analysis, different temperatures and pressures were applied to reactors R1 and R2. The minimum temperature and pressure to achieve the supercritical water was 374 °C and 25 bar, respectively, which were applied to reactor R1. For storing the syngas, it was plausible to pressurize the Reactor R2. A sensitivity analysis on the quality of the syngas and the temperature of reactor R2 was conducted to identify the optimum range of the temperature for the reactor, which was between 500 °C and 800 °C. At this range of temperature, the mole fraction of CO2 and methane was also minimized, which increased the quality of the syngas. In this research, a syngas quality (H2:CO molar ratio) greater than 2 was the target value of the simulation, which potentially has a wide range of applications in gas to liquid processes, transportation fuels, the Fischer–Tropsch process, and fuel cells. Thus, the last stage was to remove any moisture content from the syngas using a refrigerant cooler.
Figure 2 presents a schematic diagram of the water supercritical gasification, consisting of two reactors namely the supercritical gasifier (R1) and the WGS reactor (R2). To efficiently use the released heat from the exothermic reactions in the gasifier (methanation and WGS reactions), both water (stream 1), and fuel (stream 2) were fed into the supercritical gasifier (reactor R1).
To enhance the quality of syngas (H2:CO molar ratio), the gaseous products from R1 (stream 7) were fed into a WGS reactor (reactor R2). A cold water (stream 4) was fed into the reactor R2 proceeding the WGS reaction, which is exothermic and absorbs the released heat. The outlet from R2 (stream 8) was fed into a heat exchanger to produce a high-pressure steam (stream 12) from the cold water pumped into the heat exchanger (stream 3). The generated high-pressure steam was then fed into a two-stage steam turbine to produce ~11% of the total input energy to be used as electrical power for work in the plant. Then, the cold syngas from the heat exchanger (stream 10) was stored for further use. Notably, to the best of our knowledge there is no industrial WGS reaction, or any demonstration of it, which operate at the proposed conditions yet. However, the scope of this work is to thermodynamically assess the potential of the proposed system for high quality syngas production and an investigation on this subject is beyond the scope of the present investigation.
To predict the potential reactions occurring in the reactors, the Gibbs minimization method [37,38,39] was employed. To achieve this, the Aspen Plus RGibbs reactor and HSC chemistry were used to estimate the Gibbs free energy of the potential reactions. To solve the model, the following assumptions were considered:
  • The reactions reach to the equilibrium;
  • Heat loss is negligible from all reactors, pipes, tanks, and units;
  • Heat and mass transfer coefficients are plausible to maintain the conditions for highest chemical performance of the reactions;
  • Graphite in the present work is a surrogate for more realistic feedstock, which is used in the Gibbs minimization simulation. Any impurities in the feedstock have a negligible influence on the reactions and only carbon reacts in the supercritical gasifier. If there is an impurity, it is completely separated in the form of ash from the supercritical gasifier due to the difference between the ash and supercritical water density;
  • The residence time is sufficient for the reactions to reach completion so that no unreacted carbon enters the WGS reactor;
  • No catalytic effect is considered in the modelling. However, metal oxides and some composites have been identified as a suitable heat and mass transfer medium [40,41,42,43,44], which can also improve the supercritical gasification reactions [45,46,47];
  • The process is isobar and the pressure of the reactors is the same.
Table 1 shows the potential reactions occurring in the proposed concept.
To assess the chemical performance of the reactors, the following parameter is defined:
S C W C = n s t e a m n f e e d s t o c k
where, n steam is the moles of supercritical steam required for the gasification and n feedstock is the moles of feedstock fed into the reactor. SCW and C stand for supercritical water and carbon, respectively. The carbon conversion is also defined with the following equation:
x = n f e e d s t o c k , i n i t i a l n f e e d s t o c k , r e m a i n i n g n f e e d s t o c k , i n i t i a l × 100
where, n feedstock, initial is the moles of feedstock introduced to the reactor and n feedstock, remained is the unreacted moles of carbon or feedstock. To perform the sensitivity analysis and to compare the results to a reference case, a reference condition is defined in Table 2.
To validate the results of the modelling, a comparison was made between the results obtained with the developed model and those reported in the literature. To make this comparison, the model was validated using the same operating conditions given in the literature. The results of comparison showed that the estimated moles of production of hydrogen was similar to those reported in the literature [48,49,50]. As shown in Figure 3a, the moles of production of hydrogen estimated with the model is within the deviation of ±10% of those of reported in the literature [51,52]. Likewise, the syngas quality (H2:CO molar ratio), as presented in Figure 3b, was in good agreement with the literature with a deviation of ±15%.
To assess the effect of biomass composition on the composition of the gas production, three different biomasses were selected from the literature. The proximate analysis and ultimate analysis of the biomasses are listed in Table 3. It is worth noting that most of the biomass have a very low sulphur content (<1%) and a high oxygen content (>35%). The content of sulphur considerably influences the quality of the syngas. The lower the sulphur content, the higher the quality of the syngas [53,54,55].

3. Results and Discussion

3.1. Gibbs Free Energy and Enthalpy Assessment

Figure 4a presents the dependence of the Gibbs free energy on the operating temperature of the gasifier and the WGS reactor for the different reactions listed in Table 1. As shown in Figure 3a, at T > 650 °C, the Gibbs free energy for reactions 4 (methanation) and 5 (partial oxidation of graphite) is positive, meaning that these reactions are unlikely to occur at this temperature range. However, the rest of the reactions occur within the gasifier and the WGS reactors. Therefore, for reactions to be spontaneous and feasible in the gasifier, the minimum temperature should be at least 650 °C, since over this temperature, the Gibbs free energy for reactions 1, 2, 3 and 6 is negative. Figure 4b presents the dependence of enthalpy of reactions on the temperature of the gasifier and the WGS reactors for the different reactions presented in Table 1 and shows that the enthalpy of the reactions does not significantly change with temperature and remains approximately constant. For example, for each of the reactions 1, 2, 3, and 6, the enthalpy of reaction remains constant at a temperature range between 650 °C and 1000 °C. It is worth noting that the total enthalpy of the reaction for the gasifier is endothermic and it is exothermic for the WGS reactor.

3.2. Thermochemical Equilibrium Assessment for Reactor R1

Figure 5 presents the dependence of moles of gaseous products on the temperature of reactor R1 for the supercritical water to fuel ratio (SCW/C) = 1 at P = 25 bar. It is worth noting that the supercritical water gasification occurs at operating pressures larger than 25 bar. For P < 25 bar, the supercritical gasification does not occur, and the chemical conversion of carbon is very low. For an atmospheric gasification, more steam is required not only to enrich the hydrogen content but also to provide sufficient mixing in order to drive the reaction towards completion. We observe that with an increase in temperature the moles of production of H2 and CO increase. For example, at 700 °C, the mole fractions of H2 and CO are 0.86 kmol and 0.66 kmol, respectively. However, at T > 900 °C, the mole fractions o H2 and CO are 0.99 kmol and 0.97 kmol, respectively. In addition, with an increase in temperature, the moles of production for CO2 and CH4 approaches zero. Importantly, this behaviour is only seen for an SCW/C ratio of one and smaller. For example, for the SCW/C = 2, as presented in Figure 6, at 700 °C, the moles of production of H2 and CO are 1.36 kmol and 0.54 kmol, respectively. In this case, the moles of production of CO2 is 0.38 kmol, which slightly decreases with an increase in temperature. The moles of production of CO2 at 1000 °C is 0.25 kmol. Importantly, an increase in temperature slightly decreases the production of hydrogen and increases the moles of production of CO, meaning that the quality of syngas (H2:CO molar ratio) slightly decreases. Therefore, the chemical performance of the reactor R1 strongly depends on the molar ratio of SCW/C and is slightly dependent on the temperature of the reactor R1.
Figure 7 presents the dependence of the molar ratio of H2:CO (referred to as syngas quality) on the temperature of reactor R1 at the reference conditions. As shown in Figure 7, with an increase in the molar ratio of SCW/C, the quality of syngas (H2:CO molar ratio) increases, however, to achieve a target value of 2.1, a higher operating temperature is required. For example, at a molar ratio of SCW/C = 2, at T = 735 °C, the H2:CO molar ratio reached the target value, while for SCW/C = 3, the target value for the H2:CO molar ratio was satisfied at T = 988 °C. Moreover, with an increase in temperature, moles of production for CO2 decreases, hence two different regimes are seen in Figure 7. Importantly, as shown in Figure 4a, the Gibbs free energy is negative for T > 650 °C, and therefore it is likely that the reactions proceed spontaneously in this range. Therefore, these two constraints limit the operating temperature of reactor R1 between 650 °C and 900 °C. It is worth noting that at a large ratio of SCW/C, the energy requirement of the system is intensified as more steam is demanded for the supercritical reactor.
Figure 8 presents the dependence of fuel conversion on temperature for different molar ratios of SCW/C. As shown, with an increase in SCW/C, carbon conversion increases. For example, at 700 °C, for SCW/C = 0.01, the fuel conversion is only 0.8%, while for SCW/C = 1, it is 69.8%. Thus, the ratio of SCW/C strongly influences the fuel conversion. Likewise, with an increase in the temperature of the reactor, the fuel conversion increases, because with an increase in the value of SCW/C, the content of H2O increases, which drives Equations (1)–(4) including gasification, partial oxidation of carbon, methanation, and complete oxidation of carbon. These reactions cause more carbon to be consumed and the value of the carbon conversion is promoted. For example, for SCW/C = 0.5 and SCW/C = 1, at T = 800 °C the fuel conversion is 45% and 90%, respectively, however, when the temperature is increased to 1000 °C, the fuel conversion reaches 99.1% and 100%, respectively. Therefore, it is estimated that graphite is completely converted into syngas at SCW/C > 1 if the required temperature is maintained.

3.3. Thermochemical Equilibrium Assessment for Reactor R2

Figure 9 presents the dependence of mole fraction of gaseous products on the temperature of reactor R2 for different components at P = 25 bar. As shown, with an increase in temperature, the mole fraction of H2 and CO increases. For example, at T = 800 °C, the mole fractions of H2 and CO are 0.49 and 0.2, respectively, while at T = 1000 °C, they are 0.61 and 0.31, respectively. Moreover, an increase in temperature decreases the mole fractions of CO2 and CH4, respectively. For example, at 700 °C, the mole fractions of CO2 and CH4 are 0.2 and 0.18, respectively, while at 1000 °C, they are 0.04 and 0.01, respectively. Therefore, we conclude that at higher temperatures, the WGS reactor shows a better chemical performance. Notably, with an increase in the temperature of the water–gas shift reactor, the production of CO2 decreases since the water–gas shift reactor is an exothermic reactor and an increase in the temperature of the reactor decreases the chemical conversion extent of the reaction. Moreover, an increase in the temperature of the reactor causes the reaction to proceed in the reverse direction, which is endothermic. Normally, in the demonstration cases for the WGS reaction, the temperature is low (e.g., 200 °C) where the production of CO2 is maximized while the production of CO is suppressed, however, increasing the temperature results in the promotion of the production of CO2.
Figure 10 presents the dependence of mole fraction of gaseous products in syngas on the pressure of reactor R2 at T = 800 °C. According to the Le Chatelier’s principle, with an increase in pressure of the reactor, the reactions proceed towards the production of more gaseous products to reconcile the pressure effect. Therefore, increasing the pressure has no influence on the chemical performance of the WGS reactor [57]. However, pressurizing reactor R2 can maintain the pressure required for the syngas or hydrogen storage, which is advantageous. Therefore, there is no need to add a compressing unit to the plant, which reduces the cost significantly.

3.4. Biomass Composition

Figure 11a–c presents the dependence of mole fractions of gaseous products on different carbon-containing fuels.
As shown in Figure 11a, feedstock#1, has the lowest amount of carbon content and the highest amount of oxygen as compared with the other feedstocks. Therefore, more oxygen is available to produce carbon dioxide and a lower syngas quality (H2:CO molar ratio) is obtained as compared with other feedstocks. For example, at SCW/C = 1, the syngas quality (H2:CO molar ratio) is 1.02, 1.12, and 1.18, for feedstock 1, 2, and 3, respectively. It is worth noting that the low content of sulphur promotes the quality of the syngas. In addition, the lower the content of nitrogen, the higher the syngas quality (H2:CO molar ratio). Feedstock#2, as presented in Figure 11b, has the lowest amount of oxygen content, and therefore the mole fraction of CO2 is the lowest, while the quality of syngas (H2:CO molar ratio) of 2.31 is obtained at SCW/C = 2. Similarly, with an increase in the molar ratio of SCW/C, the mole fraction of H2 increases for all the biomass feedstocks, while it decreases for CO2 and CO. For example, for feedstock#1, with an increase in the SCW/C ratio from 1 to 2, the mole percentage of hydrogen increases from 35.4% to 45.2%, while the mole percentage of CO and CO2 decreases from 34.5% to 29.9% and 29.03% to 23.85%, respectively. This is attributed to the increase in the content of hydrogen and the suppression of reaction 3 due to the enrichment of hydrogen in the gas product. Since the obtained results are based on the equilibrium Gibbs minimization model, a series of experiments is required to validate and demonstrate the reactions and accurately justify the behaviour of the system based on the kinetic parameters.

4. Conclusions

A thermodynamic assessment was conducted on a supercritical water gasification with a water–gas shift reactor to assess the potential chemical performance of the proposed concept. The following conclusions were made:
  • The addition of a water–gas shift reactor adds more control on the ratio of H2:CO and minimizes the production of CH4 and CO2. For the proposed system, the quality of syngas (H2:CO molar ratio) reaches 2.1 at P = 25 bar and 850 °C and 900 °C for reactors R1 and R2, respectively.
  • Pressure was found to have no influence on the chemical performance of the water–gas shift reactor. However, pressurizing reactor R2 provides the pressure required for the storage of gas and reduces the cost of post-compression of products.
  • The proposed system produces the syngas with different H2:CO ratios depending on the SCW/C and the temperature. For a given specific biomass, we found that syngas with the quality of 2.3 was produced at 900 °C and SCW/C = 2.

Author Contributions

Conceptualization, M.M.S.; methodology, M.M.S., M.G., M.J., and M.A.; software, M.M.S. and M.J.; validation, M.M.S. and M.J.; formal analysis, M.M.S., M.J., and M.R.S.; investigation, M.M.S., M.J., M.R.S., M.G., and M.A.; writing—original draft preparation, M.M.S., M.R.S., M.J., and M.G.; writing—review and editing, M.M.S., M.J., M.R.S., M.G., and M.A.; supervision, M.A.

Funding

This research received no external funding.

Acknowledgments

The authors acknowledge the school of Mechanical Engineering, University of Adelaide for the scientific support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Turner, J.A. A realizable renewable energy future. Science 1999, 285, 687–689. [Google Scholar] [CrossRef] [PubMed]
  2. Ogden, J.M.; Williams, R.H. Solar Hydrogen: Moving Beyond Fossil Fuels; World Resources Inst: Washington, DC, USA, 1989. [Google Scholar]
  3. Klass, D.L. Biomass for Renewable Energy, Fuels, and Chemicals; Academic Press: Cambridge, MA, USA, 1998. [Google Scholar]
  4. Dry, M.E. The fischer–tropsch process: 1950–2000. Catal. Today 2002, 71, 227–241. [Google Scholar] [CrossRef]
  5. Gemmen, R.; Trembly, J. On the mechanisms and behavior of coal syngas transport and reaction within the anode of a solid oxide fuel cell. J. Power Sources 2006, 161, 1084–1095. [Google Scholar] [CrossRef]
  6. Adanez, J.; Abad, A.; Garcia-Labiano, F.; Gayan, P.; Luis, F. Progress in chemical-looping combustion and reforming technologies. Prog. Energy Combust. Sci. 2012, 38, 215–282. [Google Scholar] [CrossRef]
  7. Jafarian, M.; Arjomandi, M.; Nathan, G.J. Thermodynamic potential of high temperature chemical looping combustion with molten iron oxide as the oxygen carrier. Chem. Eng. Res. Des. 2017, 120, 69–81. [Google Scholar] [CrossRef]
  8. Jafarian, M.; Arjomandi, M.; Nathan, G.J. A hybrid solar and chemical looping combustion system for solar thermal energy storage. Appl. Energy 2013, 103, 671–678. [Google Scholar] [CrossRef]
  9. Jafarian, M.; Arjomandi, M.; Nathan, G.J. A hybrid solar chemical looping combustion system with a high solar share. Appl. Energy 2014, 126, 69–77. [Google Scholar] [CrossRef]
  10. Jafarian, M.; Arjomandi, M.; Nathan, G.J. The energetic performance of a novel hybrid solar thermal & chemical looping combustion plant. Appl. Energy 2014, 132, 74–85. [Google Scholar]
  11. Tanner, J.; Bhattacharya, S. Kinetics of CO2 and steam gasification of Victorian brown coal chars. Chem. Eng. J. 2016, 285, 331–340. [Google Scholar] [CrossRef]
  12. Patra, T.K.; Sheth, P.N. Biomass gasification models for downdraft gasifier: A state-of-the-art review. Renew. Sustain. Energy Rev. 2015, 50, 583–593. [Google Scholar] [CrossRef]
  13. Chan, F.L.; Tanksale, A. Review of recent developments in Ni-based catalysts for biomass gasification. Renew. Sustain. Energy Rev. 2014, 38, 428–438. [Google Scholar] [CrossRef]
  14. Ge, H.; Guo, W.; Shen, L.; Song, T.; Xiao, J. Biomass gasification using chemical looping in a 25kW th reactor with natural hematite as oxygen carrier. Chem. Eng. J. 2016, 286, 174–183. [Google Scholar] [CrossRef]
  15. Adánez, J.; Gayán, P.; Celaya, J.; de Diego, L.F.; García-Labiano, F.; Abad, A. Chemical looping combustion in a 10 kWth prototype using a CuO/Al2O3 oxygen carrier: Effect of operating conditions on methane combustion. Ind. Eng. Chem. Res. 2006, 45, 6075–6080. [Google Scholar] [CrossRef]
  16. Li, F.; Kim, H.R.; Sridhar, D.; Wang, F.; Zeng, L.; Chen, J.; Fan, L.-S. Syngas chemical looping gasification process: Oxygen carrier particle selection and performance. Energy Fuels 2009, 23, 4182–4189. [Google Scholar] [CrossRef]
  17. Liao, C.; Wu, C.; Yan, Y. The characteristics of inorganic elements in ashes from a 1 MW CFB biomass gasification power generation plant. Fuel Process. Technol. 2007, 88, 149–156. [Google Scholar] [CrossRef]
  18. Fan, L.; Li, F.; Ramkumar, S. Utilization of chemical looping strategy in coal gasification processes. Particuology 2008, 6, 131–142. [Google Scholar] [CrossRef]
  19. Acharya, B.; Dutta, A.; Basu, P. Chemical-looping gasification of biomass for hydrogen-enriched gas production with in-process carbon dioxide capture. Energy Fuels 2009, 23, 5077–5083. [Google Scholar] [CrossRef]
  20. Anheden, M.; Svedberg, G. Exergy analysis of chemical-looping combustion systems. Energy Convers. Manag. 1998, 39, 1967–1980. [Google Scholar] [CrossRef]
  21. Zevenhoven-Onderwater, M.; Backman, R.; Skrifvars, B.-J.; Hupa, M. The ash chemistry in fluidised bed gasification of biomass fuels. Part I: Predicting the chemistry of melting ashes and ash–bed material interaction. Fuel 2001, 80, 1489–1502. [Google Scholar] [CrossRef]
  22. Florin, N. Calcium looping technologies for gasification and reforming. In Calcium and Chemical Looping Technology for Power Generation and Carbon Dioxide (CO2) Capture; Woodhead Publishing: Sawston, UK, 2015. [Google Scholar]
  23. Matsumura, Y.; Minowa, T.; Potic, B.; Kersten, S.R.A.; Prins, W.; van Swaaij, W.P.M.; van de Beld, B.; Elliott, D.C.; Neuenschwander, G.G.; Kruse, A. Biomass gasification in near-and super-critical water: Status and prospects. Biomass Bioenergy 2005, 29, 269–292. [Google Scholar] [CrossRef]
  24. Kruse, A. Supercritical water gasification. Biofuels Bioprod. Biorefining Innov. Sustain. Econ. 2008, 2, 415–437. [Google Scholar] [CrossRef]
  25. Williams, P.T.; Onwudili, J. Subcritical and supercritical water gasification of cellulose, starch, glucose, and biomass waste. Energy Fuels 2006, 20, 1259–1265. [Google Scholar] [CrossRef]
  26. Withag, J.A.; Smeets, J.R.; Bramer, E.A.; Brem, G. System model for gasification of biomass model compounds in supercritical water–a thermodynamic analysis. J. Supercrit. Fluids 2012, 61, 157–166. [Google Scholar] [CrossRef]
  27. Guan, Q.; Savage, P.E.; Wei, C. Gasification of alga Nannochloropsis sp. in supercritical water. J. Supercrit. Fluids 2012, 61, 139–145. [Google Scholar] [CrossRef]
  28. Yakaboylu, O.; Albrecht, I.; Harinck, J.; Smit, K.; Tsalidis, G.-A.; Di Marcello, M.; Anastasakis, K.; de Jong, W. Supercritical water gasification of biomass in fluidized bed: First results and experiences obtained from TU Delft/Gensos semi-pilot scale setup. Biomass Bioenergy 2018, 111, 330–342. [Google Scholar] [CrossRef]
  29. Guo, L.; Lu, Y.; Zhang, X.; Ji, C.; Guan, Y.; Pei, A. Hydrogen production by biomass gasification in supercritical water: A systematic experimental and analytical study. Catal. Today 2007, 129, 275–286. [Google Scholar] [CrossRef]
  30. Yu, D.; Aihara, M.; Antal, M.J., Jr. Hydrogen production by steam reforming glucose in supercritical water. Energy Fuels 1993, 7, 574–577. [Google Scholar] [CrossRef]
  31. Yong, T.L.-K.; Matsumura, Y. Reaction kinetics of the lignin conversion in supercritical water. Ind. Eng. Chem. Res. 2012, 51, 11975–11988. [Google Scholar] [CrossRef]
  32. Zhu, C.; Wang, R.; Jin, H.; Lian, X.; Guo, L.; Huang, J. Supercritical water gasification of glycerol and glucose in different reactors: The effect of metal wall. Int. J. Hydrog. Energy 2016, 41, 16002–16008. [Google Scholar] [CrossRef]
  33. Reddy, S.N.; Nanda, S.; Dalai, A.K.; Kozinski, J.A. Supercritical water gasification of biomass for hydrogen production. Int. J. Hydrogen Energy 2014, 39, 6912–6926. [Google Scholar] [CrossRef]
  34. Tang, H.; Kitagawa, K. Supercritical water gasification of biomass: Thermodynamic analysis with direct Gibbs free energy minimization. Chem. Eng. J. 2005, 106, 261–267. [Google Scholar] [CrossRef]
  35. Castello, D.; Fiori, L. Supercritical water gasification of biomass: Thermodynamic constraints. Bioresour. Technol. 2011, 102, 7574–7582. [Google Scholar] [CrossRef] [PubMed]
  36. Lu, Y.; Zhao, L.; Han, Q.; Wei, L.; Zhang, X.; Guo, L.; Wei, J. Minimum fluidization velocities for supercritical water fluidized bed within the range of 633–693 K and 23–27 MPa. Int. J. Multiph. Flow 2013, 49, 78–82. [Google Scholar] [CrossRef]
  37. Sarafraz, M.M.; Jafarian, M.; Arjomandi, M.; Nathan, G.J. The relative performance of alternative oxygen carriers for liquid chemical looping combustion and gasification. Int. J. Hydrogen Energy 2017, 42, 16396–16407. [Google Scholar] [CrossRef]
  38. Sarafraz, M.M.; Jafarian, M.; Arjomandi, M.; Nathan, G.J. Potential of molten lead oxide for liquid chemical looping gasification (LCLG): A thermochemical analysis. Int. J. Hydrogen Energy 2018, 43, 4195–4210. [Google Scholar] [CrossRef]
  39. Sarafraz, M.M.; Jafarian, M.; Arjomandi, M.; Nathan, G.J. The thermo-chemical potential liquid chemical looping gasification with bismuth oxide. Int. J. Hydrogen Energy 2019, 44, 8038–8050. [Google Scholar] [CrossRef]
  40. Salari, E.; Peyghambarzadeh, M.; Sarafraz, M.M.; Hormozi, F. Boiling heat transfer of alumina nano-fluids: Role of nanoparticle deposition on the boiling heat transfer coefficient. Period. Polytech. Chem. Eng. 2016, 60, 252–258. [Google Scholar] [CrossRef]
  41. Salari, E.; Peyghambarzadeh, S.M.; Sarafraz, M.M.; Hormozi, F.; Nikkhah, V. Thermal behavior of aqueous iron oxide nano-fluid as a coolant on a flat disc heater under the pool boiling condition. Heat Mass Transf. 2017, 53, 265–275. [Google Scholar] [CrossRef]
  42. Sarafraz, M.M.; Arya, A.; Nikkhah, V.; Hormozi, F. Thermal performance and viscosity of biologically produced silver/coconut oil nanofluids. Chem. Biochem. Eng. Q. 2017, 30, 489–500. [Google Scholar] [CrossRef]
  43. Arya, A.; Sarafraz, M.M.; Shahmiri, S.; Madani, S.A.H.; Nikkhah, V.; Nakhjavani, S.M. Thermal performance analysis of a flat heat pipe working with carbon nanotube-water nanofluid for cooling of a high heat flux heater. Heat Mass Transf. 2018, 54, 985–997. [Google Scholar] [CrossRef]
  44. Sarafraz, M.M.; Arjomandi, M. Demonstration of plausible application of gallium nano-suspension in microchannel solar thermal receiver: Experimental assessment of thermo-hydraulic performance of microchannel. Int. Commun. Heat Mass Transf. 2018, 94, 39–46. [Google Scholar] [CrossRef]
  45. Guo, Y.; Wang, S.Z.; Xu, D.H.; Gong, Y.M.; Ma, H.H.; Tang, X.Y. Review of catalytic supercritical water gasification for hydrogen production from biomass. Renew. Sustain. Energy Rev. 2010, 14, 334–343. [Google Scholar] [CrossRef]
  46. Chakinala, A.G.; Brilman, D.W.F.; van Swaaij, W.P.M.; Kersten, S.R.A. Catalytic and non-catalytic supercritical water gasification of microalgae and glycerol. Ind. Eng. Chem. Res. 2009, 49, 1113–1122. [Google Scholar] [CrossRef]
  47. Yanik, J.; Ebale, S.; Kruse, A.; Saglam, M.; Yüksel, M. Biomass gasification in supercritical water: II. Effect of catalyst. Int. J. Hydrogen Energy 2008, 33, 4520–4526. [Google Scholar] [CrossRef]
  48. Aziz, M. Integrated supercritical water gasification and a combined cycle for microalgal utilization. Energy Convers. Manag. 2015, 91, 140–148. [Google Scholar] [CrossRef]
  49. Wan, W. An innovative system by integrating the gasification unit with the supercritical water unit to produce clean syngas: Effects of operating parameters. Int. J. Hydrogen Energy 2016, 41, 14573–14582. [Google Scholar] [CrossRef]
  50. Wan, W. An innovative system by integrating the gasification unit with the supercritical water unit to produce clean syngas for solid oxide fuel cell (SOFC): System performance assessment. Int. J. Hydrogen Energy 2016, 41, 22698–22710. [Google Scholar] [CrossRef]
  51. Li, N.; Li, Y.; Ban, Y.; Song, Y.; Zhi, K.; Teng, Y.; He, R.; Zhou, H.; Liu, Q.; Qi, Y. Direct production of high hydrogen syngas by steam gasification of Shengli lignite/chars: Remarkable promotion effect of inherent minerals and pyrolysis temperature. Int. J. Hydrogen Energy 2017, 42, 5865–5872. [Google Scholar] [CrossRef]
  52. Calzavara, Y.; Joussot-Dubien, C.; Boissonnet, G.; Sarrade, S. Evaluation of biomass gasification in supercritical water process for hydrogen production. Energy Convers. Manag. 2005, 46, 615–631. [Google Scholar] [CrossRef]
  53. McKendry, P. Energy production from biomass (part 1): Overview of biomass. Bioresour. Technol. 2002, 83, 37–46. [Google Scholar] [CrossRef]
  54. Indrawan, N.; Thapa, S.; Bhoi, P.R.; Huhnke, R.L.; Kumar, A. Engine power generation and emission performance of syngas generated from low-density biomass. Energy Convers. Manag. 2017, 148, 593–603. [Google Scholar] [CrossRef]
  55. Ong, Z.; Cheng, Y.; Maneerung, T.; Yao, Z.; Tong, Y.W.; Wang, C.H.; Dai, Y. Co-gasification of woody biomass and sewage sludge in a fixed-bed downdraft gasifier. AIChE J. 2015, 61, 2508–2521. [Google Scholar] [CrossRef]
  56. Zhai, Y.; Peng, C.; Xu, B.; Wang, T.; Li, C.; Zeng, G.; Zhu, Y. Hydrothermal carbonisation of sewage sludge for char production with different waste biomass: Effects of reaction temperature and energy recycling. Energy 2017, 127, 167–174. [Google Scholar] [CrossRef]
  57. Bustamante, F.; Enick, R.; Rothenberger, K.; Howard, B.; Cugini, A.; Ciocco, M.; Morreale, B. Kinetic study of the reverse water gas shift reaction in high-temperature, high pressure homogeneous systems. Fuel Chem. Div. Prepr. 2002, 47, 663–664. [Google Scholar]
Figure 1. Detailed operating conditions of the process for supercritical synthetic gas (syngas) production.
Figure 1. Detailed operating conditions of the process for supercritical synthetic gas (syngas) production.
Energies 12 02591 g001
Figure 2. Schematic diagram of the supercritical gasification.
Figure 2. Schematic diagram of the supercritical gasification.
Energies 12 02591 g002
Figure 3. Validation of the model using the data reported in the literature, (a) comparison between estimated moles of production of hydrogen and data reported in previous works [48,49,50], (b) comparison between the estimated syngas quality (H2:CO molar ratio) and those reported in the literature [51,52].
Figure 3. Validation of the model using the data reported in the literature, (a) comparison between estimated moles of production of hydrogen and data reported in previous works [48,49,50], (b) comparison between the estimated syngas quality (H2:CO molar ratio) and those reported in the literature [51,52].
Energies 12 02591 g003
Figure 4. Dependence on temperature of the Gibbs free energy and enthalpy of the reaction with temperature. (a) Variation of the change in the Gibbs free energy of reactions with temperature for the reactions given in Table 1 and (b) variation of the enthalpy of the reaction with temperature for the reactions given in Table 1.
Figure 4. Dependence on temperature of the Gibbs free energy and enthalpy of the reaction with temperature. (a) Variation of the change in the Gibbs free energy of reactions with temperature for the reactions given in Table 1 and (b) variation of the enthalpy of the reaction with temperature for the reactions given in Table 1.
Energies 12 02591 g004
Figure 5. Variation of the mole fraction of the productions with temperature of the reactor R1 at reference conditions given in Table 2.
Figure 5. Variation of the mole fraction of the productions with temperature of the reactor R1 at reference conditions given in Table 2.
Energies 12 02591 g005
Figure 6. Variation of the mole fraction of the products with temperature at the reference condition in reactor at the supercritical water to fuel ratio (SCW/C) = 2.
Figure 6. Variation of the mole fraction of the products with temperature at the reference condition in reactor at the supercritical water to fuel ratio (SCW/C) = 2.
Energies 12 02591 g006
Figure 7. Variation of the syngas quality (H2:CO molar ratio) with temperature, for different values of the supercritical water to fuel ratio (SCW/C) in reactor R1. The target (shown in the figure) is to obtain the syngas quality (H2: CO molar ratio) value of ~2.1, suitable for the Fischer–Tropsch process and some fuel cell applications.
Figure 7. Variation of the syngas quality (H2:CO molar ratio) with temperature, for different values of the supercritical water to fuel ratio (SCW/C) in reactor R1. The target (shown in the figure) is to obtain the syngas quality (H2: CO molar ratio) value of ~2.1, suitable for the Fischer–Tropsch process and some fuel cell applications.
Energies 12 02591 g007
Figure 8. Variation of the carbon (fuel) conversion extent with temperature, for various values of SCW/C ratio in reactor R1.
Figure 8. Variation of the carbon (fuel) conversion extent with temperature, for various values of SCW/C ratio in reactor R1.
Energies 12 02591 g008
Figure 9. Variation of the mole fraction of the product with temperature at the reference condition for reactor R2.
Figure 9. Variation of the mole fraction of the product with temperature at the reference condition for reactor R2.
Energies 12 02591 g009
Figure 10. Variation of the mole fraction of the products with pressure at the reference condition in reactor R2.
Figure 10. Variation of the mole fraction of the products with pressure at the reference condition in reactor R2.
Energies 12 02591 g010
Figure 11. Mole percentage of components in gaseous products for different molar ratios of SCW/C and for different biomass feedstock given in Table 2 and at 900 °C (a) feedstock #1, (b) feedstock #2, and (c) feedstock #3.
Figure 11. Mole percentage of components in gaseous products for different molar ratios of SCW/C and for different biomass feedstock given in Table 2 and at 900 °C (a) feedstock #1, (b) feedstock #2, and (c) feedstock #3.
Energies 12 02591 g011
Table 1. Potential reactions of supercritical water with carbon.
Table 1. Potential reactions of supercritical water with carbon.
No.NameReaction
1ReformingC + H2O(g) Energies 12 02591 i001 H2(g) + CO(g)
2Partial oxidation of graphite3C + 2H2O(g) Energies 12 02591 i001 CH4(g) + 2CO(g)
3Complete oxidation of graphiteC + 2H2O(g) Energies 12 02591 i001 2H2(g) + CO2(g)
4MethanationC + 2H2(g) Energies 12 02591 i001 CH4(g)
5Water-gas shift reactionCO(g) + H2O(g) Energies 12 02591 i001 H2(g) + CO2(g)
6Boudouard reactionCO2(g) + C Energies 12 02591 i001 2CO(g)
Table 2. Reference conditions used in the sensitivity analysis.
Table 2. Reference conditions used in the sensitivity analysis.
ParameterT R1 (°C)T R2 (°C)Pressure (bar)SCW/C
Range (min–max)650–1000500–80025–600.01–3
Reference case650600251
Table 3. Analysis of the agro biomass feedstock used in this research [56].
Table 3. Analysis of the agro biomass feedstock used in this research [56].
FeedstockProximate AnalysisUltimate Analysis
BiomassMoisture ContentAshVolatile MatterFixed CarbonCHNSO
Feedstock 155.3160.3230.329.3622.33.32.10.511.3
Feedstock 247.9840.2434.5525.2150.23.82.70.52.4
Feedstock 341.3952.6732.1915.1437.83.12.30.423.6

Share and Cite

MDPI and ACS Style

Sarafraz, M.M.; Safaei, M.R.; Jafarian, M.; Goodarzi, M.; Arjomandi, M. High Quality Syngas Production with Supercritical Biomass Gasification Integrated with a Water–Gas Shift Reactor. Energies 2019, 12, 2591. https://doi.org/10.3390/en12132591

AMA Style

Sarafraz MM, Safaei MR, Jafarian M, Goodarzi M, Arjomandi M. High Quality Syngas Production with Supercritical Biomass Gasification Integrated with a Water–Gas Shift Reactor. Energies. 2019; 12(13):2591. https://doi.org/10.3390/en12132591

Chicago/Turabian Style

Sarafraz, M. M., Mohammad Reza Safaei, M. Jafarian, Marjan Goodarzi, and M. Arjomandi. 2019. "High Quality Syngas Production with Supercritical Biomass Gasification Integrated with a Water–Gas Shift Reactor" Energies 12, no. 13: 2591. https://doi.org/10.3390/en12132591

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

Article Metrics

Back to TopTop