Next Article in Journal
The Climate-Independent Need for Renewable Energy in the 21st Century
Next Article in Special Issue
Biodiesel Production Potential from Littered Edible Oil Fraction Using Directly Synthesized S-TiO2/MCM-41 Catalyst in Esterification Process via Non-Catalytic Subcritical Hydrolysis
Previous Article in Journal
Icing Forecasting of Transmission Lines with a Modified Back Propagation Neural Network-Support Vector Machine-Extreme Learning Machine with Kernel (BPNN-SVM-KELM) Based on the Variance-Covariance Weight Determination Method
Previous Article in Special Issue
Influence of Catalytic Formulation and Operative Conditions on Coke Deposition over CeO2-SiO2 Based Catalysts for Ethanol Reforming
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Combustion and Heat Release Characteristics of Biogas under Hydrogen- and Oxygen-Enriched Condition

1
Department of Chemical Engineering, Nagoya University, Nagoya, Aichi 464-8603, Japan
2
Guangzhou Institute of Energy Conversion, Chinese Academy of Sciences, Guangzhou 510640, China
3
Faculty of Mechanical Engineering, Kanazawa University, Kakuma, Kanazawa, Ishikawa 920-1192, Japan
*
Authors to whom correspondence should be addressed.
Energies 2017, 10(8), 1200; https://doi.org/10.3390/en10081200
Submission received: 10 May 2017 / Revised: 5 July 2017 / Accepted: 20 July 2017 / Published: 13 August 2017
(This article belongs to the Collection Bioenergy and Biofuel)

Abstract

:
Combustion and heat release characteristics of biogas non-premixed flames under various hydrogen-enriched and oxygen-enriched conditions were investigated through chemical kinetics simulation using detailed chemical mechanisms. The heat release rates, chemical reaction rates, and molar fraction of all species of biogas at various methane contents (35.3–58.7%, mass fraction), hydrogen addition ratios (10–50%), and oxygen enrichment levels (21–35%) were calculated considering the GRI 3.0 mechanism and P1 radiation model. Results showed that the net reaction rate of biogas increases with increasing hydrogen addition ratio and oxygen levels, leading to a higher net heat release rate of biogas flame. Meanwhile, flame length was shortened with the increase in hydrogen addition ratio and oxygen levels. The formation of free radicals, such as H, O, and OH, are enhanced with increase in hydrogen addition ratio and oxygen levels. Higher reaction rates of exothermic elementary reactions, especially those with OH free radical are increased, are beneficial to the improvement in combustion and heat release characteristics of biogas in practical applications.

1. Introduction

With continuous depletion of fossil fuels and worsening pollution, research and development of renewable energy sources have attracted considerable interest [1,2,3,4,5,6,7]. Biogas can be produced using an anaerobic digester of biomass or biodegradable organic wastes. Compositions of biogas vary with the feedstock and the fermentation process [1,2,3]. The main compositions of biogas are methane (CH4) and carbon dioxide (CO2), as well as a small amount of water (H2O), nitrogen (N2), and hydrogen (H2) [3,4,5].
As an advantageous energy source, biogas is widely considered as a suitable fuel for heating facilities, power generation, and vehicles [8,9,10,11,12,13,14,15]. However, biogas possesses a relatively lower heating value (about 3000–6000 kcal·m−3) and lower heat release rate than conventional fuels, which limits the application of biogas in combustion devices. Somehsaraei et al. [16] developed a steady state thermodynamic model for biogas application in 100 kW micro gas turbines. They found that mass flow and pressure ratio in the micro gas turbines are lower than those of the natural gas–fueled turbines. Furthermore, the total electrical efficiency decreases with decreasing CH4 content in biogas fuel, and the use of biogas negatively affects heat recovery. A three-dimensional computational fluid dynamic study on biogas flameless mode was developed by Hosseini et al. [17]. Although increasing preheated temperature positively affects the reduction of fuel consumption, a fuel with high calorific value is necessary to preheat the furnace prior to the application of biogas combustion in the combustion furnace. The flameless combustion temperature of a biogas furnace is lower than traditional combustion throughout a combustion chamber. Several experimental and numerical studies have explored the fuel characteristics of biogas. In such studies, high-quality fuel such as H2 [18, 19, 20, 21, 22] is blended with biogas to improve the relatively poor combustion characteristics of biogas. Zhen et al. [18] experimentally investigated the effect of H2 addition on the characteristics of a biogas diffusion flame. They found that H2 addition significantly improved the stability of the biogas flame. The flame temperature increased and visible flame length reduced with increasing H2 addition ratios. Wei et al. [19] numerically investigated the effects of equivalence ratio, H2, and CO2 on the heat release characteristics of the premixed laminar biogas-hydrogen flames. The results showed that OH + H2 ⇔ H + H2O was major endothermic reaction for biogas-hydrogen premixed flames. The total heat release was enhanced evidently with H2 addition. Non-premixed are widely applied in the industrial process, while the heat release rate of biogas flame is low, which obstructs the application of biogas in practical application. Therefore, the improvement of heat release characteristics of biogas flame at various condition are necessary.
The above results show that numerous studies have investigated the combustion characteristics of biogas premixed flame. Nevertheless, understanding the non-premixed combustion of a fuel in practical application is necessary. To utilize biogas in internal combustion engines, the unburnt composition CO2 or N2 is expected to be excluded from biogas. Therefore, mechanical biological treatment is needed to upgrade raw biogas and increase CH4 content in biogas. The high cost of such treatment process is unfavorable to the application of biogas as a fuel. Therefore, the direct application of raw biogas is more economic for engines or heat supply. The current research proposes biogas combustion under hydrogen-enriched and oxygen-enriched conditions as potential methods to improve biogas combustion. The non-premixed flames of biogas under various hydrogen-enriched and oxygen-enriched conditions are studied, and the combustion and heat release characteristics of biogas non-premixed flames are numerically investigated. A non-premixed combustion model of biogas is established. The GRI 3.0 mechanism with 53 species and 325 elementary chemical reactions is employed, and the P1 radiation model is considered. All elementary chemical reaction rates and heat release rates are programmed. To enhance the combustion and heat release characteristics, hydrogen addition and oxygen enriched combustion are selected for biogas combustion in future practical application. The mass contents of CH4 in biogas, hydrogen addition ratios, and oxygen enrichment levels are ranged from 35.3% to 58.7%, 0% to 50%, and 21% to 35%, respectively.

2. Numerical Simulations

A typical gaseous combustion cylindrical chamber with a diameter of 200 mm (r) and 600 mm length (z) is employed in this research, as shown in Figure 1. The fuel inlet is located at a central nozzle with a radius of 6.8 mm, and air inlet is surrounded with a fuel inlet with a radius of 100 mm. Three kinds of biogas fuel are applied in this research. The H2 addition ratio to fuel ( X H 2 , based on low heat value) and O2-enriched level ( Ω O 2 ) are defined as following equations:
X H 2   =   V H 2   ×   L H V H 2 V CH 4   ×   L H V CH 4   +   V H 2 × L H V H 2
Ω O 2   =   V O 2 V O 2   +   V N 2
where V CH 4 , V H 2 , V O 2 , and V N 2 are the flow rates of CH4, H2, O2, and N2 respectively; L H V H 2 and L H V CH 4 are the low heat values of H2 and CH4. The inlet fuel and oxidizer conditions at various H2-enriched condition and O2-enriched condition are shown in Table 1. The inlet power is kept constant under all calculation conditions. The mass contents of CH4 in biogas are 35.3%, 45.9%, and 58.7%, namely BG-1, BG-2, and BG-3, respectively. H2 addition ratio to biogas is arranged from 10% to 50%, and O2 in the oxidizer is from 21% to 35%.
The flow mass, chemical species, momentum, and energy equation based on Navier–Stokes equations is expressed in Equation (3):
δ ( ρ Y ) δ t   +   d i v ( ρ V Y )   =   d i v ( Γ Y g r a d Y )   +   S Y
where ρ , Y , V , Γ Y , and S Y are density, general dependent variable, velocity vector, associated transport coefficient of Y, and the source term. Thermal radiation is programmed to the source term of the mixture enthalpy. In this research, P1 radiation developed by Grosshandler of National Institute of Standards and Technology (NIST) [23] was selected. The gas radiation was considered by the determination of Planck mean absorption coefficient. Gas radiation contributed to the gases of CH4, CO2, CO, and H2O. The mean absorption coefficient of these gases ( α i ) can be fitted as functions of temperature, which are summarized as Equations (4)–(6):
Mean absorption coefficient of CH4:
α i   =   a 0   +   a 1 T   +   a 2 T 2   +   a 3 T 3   +   a 4 T 4   +   a 5 T 5
Mean absorption coefficient of CO2 and H2O:
α i = a 0   +   a 1 ( 1000 T )   +   a 2 ( 1000 T ) 2   +   a 3 ( 1000 T ) 3   +   a 4 ( 1000 T ) 4   +   a 5 ( 1000 T ) 5
Mean absorption coefficient of CO:
α i   =   a 0   +   T { a 1   +   T [ a 2   +   T ( a 3   +   a 4 T ) ] }
The values of constants a 0 to a 5 for CH4, CO2, CO, and H2O are shown in Table 2. The boundary wall is treated as a gray heat sink with emissivity of 0.8, and the wall temperature is maintained at 300 K through water cooling. The total absorption coefficient of radiating gas mixture is calculated as the sum of the absorption coefficients of CH4, CO2, CO, and H2O in the flame. All transport equations with radiation are solved by CFD software PHOENICS 2014 (Cham-Japan, Tokyo, Japan) [24]. The velocity and pressure coupling are treated by a SIMPLE method algorithm [25]. The GRI 3.0 mechanism with 53 species and 325 reactions [26] is applied to this research. The transport and thermal properties of all species are calculated by GRI 3.0 database and CHEMKIN codes. As a key parameter in fuel combustion, the heat release rates of all elementary reactions are defined and programed. Taking reaction R52 as example, we can calculate the heat release rate by following equations:
R 52 : H   +   CH 3 ( + M )     CH 4 ( + M )
Δ h i = 52   =   ( h H + h CH 3 )     h CH 4 ,   J · mol 1
q i = 52   =   Δ h i = 52   ×   k i = 52
q net   =   i = 1 325 q i
where h j is formation enthalpy of specie j (J·mol−1); Δ h i , k i , and q i are enthalpy change of reaction i (J·mol−1), the reaction rate of reaction i (mol·cm−3·s−1), and heat release of reaction i (J·cm−3·s−1), respectively. Furthermore, q net indicates the total heat release rate of all reactions (mol·cm−3·s−1).

3. Discussion

Figure 2 shows several typical measured outlet temperatures of BG-2/air flames with various H2 addition ratios and O2-enriched levels at the same input power of 1.0 kW and equivalence ratio of 0.95; the simulation results are shown for comparison. The outlet temperature was measured by a K-type thermocouple on a same combustion chamber and burner size with simulation, which was described in our previous study [27]. The error of experimental temperature at all H2 addition ratios and O2-eneriched levels are less than 15%. Furthermore, the deviation of simulation and experimental results are below 11%. Therefore, the simulation results show good agreement with experimental data at all H2 addition ratios and O2-enriched levels, thereby confirming that the GRI 3.0 mechanism provides an acceptable and repeatable result for biogas combustion under H2-enriched and O2-enriched conditions.
The effect of CH4 content in biogas fuel on the combustion temperature distribution is shown in Figure 3. The maximum combustion temperature increases with increasing CH4 content in biogas fuel. Furthermore, the flame length through the z-direction decreases at a higher CH4 content condition, and flame width through the r-direction also shrinks with increasing CH4 content in biogas fuel. These findings are attributed to the following explanations: (a) the reaction rate of elementary reactions in biogas fuel flames increases with increasing CH4 content in biogas fuel as shown in Figure 4, and is usually related to the flame structure; (b) the heat loss to heat up unburnt CO2 in biogas fuel gas decreases with higher CH4 content in biogas fuel; and (c) the total volume of combustion gas (including fuel gas and air) decreases when CH4 content in biogas fuel increases, leading to higher temperature and heat release rate under higher CH4 content condition. The maximum heat release rate increased by about 19.3% (from 362 to 432 J·cm−3·s−1) with increasing CH4 content in biogas from BG-1 to BG-3, leading to a stronger heat release rate flame with shorter flame lengths at higher CH4 content in biogas fuel.
The effects of H2 addition ratios and O2 enrichment levels in BG-2/air flame on combustion temperature distributions are shown in Figure 5. The maximum combustion temperature in flames increased by approximately 10.8% from 1803 to 1998 K as H2 addition ratio varied from 0% to 50%, and was slightly higher by about 11.2% from 1803 to 2005 K as O2 enrichment levels increased from 21% to 35%. Conversely, the maximum heat release rate in flames increased by 3.69 times from 362 to 1336 J·cm−3·s−1 at H2 addition conditions and only 2.29 times from 362 to 829 J·cm−3·s−1 under O2 addition conditions, as shown in Figure 6. This difference is also related to the shape and structure of biogas flames, which is important in the direct application of biogas for heating. Specifically, the position of maximum heat release rate varying with H2 addition and O2 enrichment is different. At the H2 addition condition, the fuel composition at fuel inlet varies with H2 addition rate, the fuel combustion velocity enhances because of high reactivity (more OH and H radicals with higher formation rate because of high burning velocity [28,29,30] and combustion temperature of H2) and high mobility of H2 [28,29,30]. Therefore, the maximum heat release rate at the H2 addition conditions is located at center of the flame. At the O2 enrichment condition, the fuel composition keeps constant, and the air composition at outside of fuel inlet varies with O2 enrichment. The fuel combustion velocity also improves because of the higher collision rate of oxidizer with fuel near flame edges. Therefore, the maximum heat release rate locates near the flame edges at O2 enriched condition.
To obtain an in-depth understanding of heat release characteristics of biogas flames under H2-enriched and O2-enriched conditions, the net heat release rate distribution at centerline along the z-direction is shown in Figure 7. The maximum net heat release rate at centerline is enhanced by approximately two orders of magnitude (from 0.155 to 120.0 J·cm−3·s−1) when H2 addition ratio increases from 0% to 50%; and the maximum net heat release rate at centerline is enhanced by about one order of magnitude (from 0.155 to 3.66 J·cm−3·s−1) when the O2 addition ratio increases from 21% to 35%. The enhancement of the heat release rate can lead to a change in flame structure and heat release characteristics in the flame reaction zone. Therefore, average temperature Ta was defined to explore the effect of H2-enriched and O2-enriched conditions on flame kernel. Ta is based on the barycenter of heat release rate as follows:
T a   =   0 z q T d z 0 z q d z
Figure 8 shows the average temperature Ta together with flame height of peak heat release rate at various H2 addition ratios and O2 enrichment levels. As shown in Figure 8, the average temperature Ta shows a gradual decrease with the increasing H2 addition ratio and O2 enrichment levels. For the biogas flame at same input power of 1 kW and equivalence ratio of 0.95, the value of average temperature Ta increases from 1215 to 1864 K, nearly yielding an enhancement of 53.4%, when H2 addition ratio varied from 0% to 50%. In contrast, the flame height of peak heat release rate decreases by approximately 54.5% from 19.8 to 9.0 cm when the H2 addition ratio increases from 0% to 50%. The enhancement of average temperature Ta and reduction of flame height of maximum heat release rate resulted from the high mobility and high reactivity of H2, leading to a higher reaction rate of biogas flame under H2-enriched condition as shown in Figure 9a. By contrast, the value of average temperature Ta increased from 1215 to 1428 K (~17.5% increase) and the flame height of peak heat release rate decreases from 19.8 to 10.9 cm (~44.9% decrease) when O2 enrichment levels increased from 21% to 35%. With the enrichment of O2 in combustion air, the reaction rate can be improved as shown in Figure 9b. Furthermore, unnecessary sensible heat to heat up the unburnt N2 gas also decreases with increasing O2 enrichment levels. As shown in Figure 9, the net reaction rate of biogas flames under the H2-enriched condition markedly increases as H2 addition ratio increases from 0% to 50% and lower increase is observed under O2-enriched condition when O2 levels increases from 21% to 35%. To clarify the main elementary reaction in biogas flame, we compared the typical elementary exothermic reaction rate of biogas flame along the centerline at H2 addition ratio of 0% and 50% in Figure 10. The formation of free radicals, such as H, O, and OH, are enhanced with the increase in hydrogen addition ratios; in particular, R83: OH + H2 ⇔ H + H2O markedly increases as H2 is added to biogas flame, similar results have been reported for the heat release rate of premixed laminar biogas-H2 flame by Wei et al. [19] and CH4/air premixed flames with H2 addition by Hu et al. [31]. Furthermore, the typical elementary exothermic reaction rates of R83: OH + H2 ⇔ H + H2O; R37: H + O2 ⇔ O + OH are important to enhance biogas combustion. The elementary reaction of R83 (H2 as reactant) is higher than R37 (O2 as reactant). Therefore, H2 addition has a drastic effect on heat release rate (and reaction rate) enhancement than O2 enrichment as shown in Figure 6 and Figure 7. The enhancement reaction rate can lead to structure and shape changes in biogas flame. The flame thickness along centerline under H2-enriched conditions and O2-enriched conditions are shown in Figure 11. The definition of flame thickness can be expressed as follows:
D F   =   T m a x T 0 ( d T / d z ) m a x
where T m a x and T 0 are the maximum temperature and unburned temperature respectively, ( d T / d z ) m a x is the peak gradient in temperature profile along the centerline. As shown in Figure 11, the flame thickness decreases with the increasing H2 addition ratio and O2 enrichment levels. This decrease can be attributed to the burning velocity enhancement of biogas flames under H2-enriched and O2-enriched conditions. Therefore, a flame with enhanced heat release rate is formed under H2-enriched and O2-enriched conditions, which are beneficial to the practical application of biogas as a fuel.

4. Conclusions

The combustion and heat release characteristics of biogas non-premixed flames under various hydrogen-enriched and oxygen-enriched conditions were investigated utilizing chemical kinetics simulation and the detailed chemical mechanism. The results are summarized as follows.
(1)
The net reaction rate of biogas increases with increasing hydrogen addition ratio and oxygen levels, leading to a higher net heat release rate of biogas flame;
(2)
The formation of free radicals, such as H, O, and particularly OH, are enhanced with the increase in hydrogen addition ratio and oxygen levels;
(3)
Flames with enhanced heat release rates are formed under H2-enriched and O2-enriched conditions. Therefore, H2-enriched and O2-enriched combustion is beneficial to the improvement of combustion and heat release characteristics of biogas in practical application.

Acknowledgments

This research is partially supported by the “Knowledge Hub Aichi” Priority Research Project from Aichi Prefectural Government, Japan.

Author Contributions

Jun Li contributed to the experiment and simulation, data analysis and paper writing; Hongyu Huang, Huhetaoli, Yu Bai, and Yong Chen contributed to experimental design, and Yugo Osaka and Noriyuki Kobayashi contributed to modeling selection and programing instruction.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

X H 2 H2 addition ratio to fuel (based on low heat value)
V H 2 Flow rate of H2 (m·s−1)
V CH 4 Flow rate of CH4 (m·s−1)
L H V H 2 Low heat value of H2 (kJ·mol−1)
L H V CH 4 Low heat value of CH4 (kJ·mol−1)
Ω O 2 O2-enriched level (-)
V O 2 Flow rate of O2 (m·s−1)
V N 2 Flow rate of N2 (m·s−1)
ρ Density (kg·m−3)
Y General dependent variable (-)
Y Velocity vector (m·s−1)
Γ Y associated transport coefficient of Y (-)
S Y The source term (-)
α i Mean absorption coefficient of species i (-)
h j Formation enthalpy of specie j (J·mol−1)
Δ h i Enthalpy change of reaction i (J·mol−1)
k i The reaction rate of reaction i (mol·cm−3·s−1)
q i Heat release rate of reaction i (J·cm−3·s−1)
q net Total heat release rate of all reaction (J·cm−3·s−1)
D F Flame thickness (cm)
T m a x The maximum temperature (K)
T 0 The unburned temperature (K)
( d T / d z ) m a x The peak gradient in temperature (K·cm−1)

References

  1. Divya, D.; Gopinath, L.R.; Christy, P.M. A review on current aspects and diverse prospects for enhancing biogas production in sustainable means. Renew. Sustain. Energy Rev. 2015, 42, 690–699. [Google Scholar] [CrossRef]
  2. Pazera, A.; Slezak, R.; Krzystek, L.; Ledakowicz, S.; Bochmann, G.; Gabauer, W.; Helm, S.; Reitmeier, S.; Marley, L.; Gorga, F.; et al. Biogas in Europe: Food and Beverage (FAB) Waste Potential for Biogas Production. Energy Fuels 2015, 29, 4011–4021. [Google Scholar] [CrossRef]
  3. Papurello, D.; Soukoulis, C.; Schuhfried, E.; Cappellin, L.; Gasperi, F.; Silvestri, S.; Santarelli, M.; Biasioli, F. Monitoring of volatile compound emissions during dry anaerobic digestion of the organic fraction of municipal solid waste by proton transfer reaction time-of-flight mass spectrometry. Bioresour. Technol. 2012, 126, 245–265. [Google Scholar] [CrossRef] [PubMed]
  4. Papurello, D.; Lanzini, A.; Tognana, L.; Silvestri, S.; Santarelli, M. Waste to energy: Exploitation of biogas from organic waste in a 500 Wel solid oxide fuel cell (SOFC) stack. Energy 2015, 85, 145–158. [Google Scholar] [CrossRef]
  5. Papurello, D.; Silvestri, S.; Tomasi, L.; Belcari, I.; Biasioli, F.; Santarelli, M. Biowaste for SOFCs. Energy Procedia 2016, 101, 424–431. [Google Scholar] [CrossRef]
  6. Bensaid, S.; Russo, N.; Fino, D. Power and hydrogen co-generation from biogas. Energy Fuels 2010, 24, 4743–4747. [Google Scholar] [CrossRef]
  7. Wang, R.; Zhao, Z.; Liu, J.; Lv, Y.; Ye, X. Enhancing the storage stability of petroleum coke slurry by producing biogas from sludge fermentation. Energy 2016, 113, 319–327. [Google Scholar] [CrossRef]
  8. Wang, J.; Xie, Y.; Cai, X.; Nie, Y.; Peng, C.; Huang, Z. Effect of H2O addition on the flame front evolution of syngas spherical propagation flames. Combust. Sci. Technol. 2016, 188, 1054–1072. [Google Scholar] [CrossRef]
  9. Park, C.; Park, S.; Lee, Y.; Kim, C.; Lee, S.; Moriyoshi, Y. Performance and emission characteristics of a SI engine fueled by low calorific biogas blended with hydrogen. Int. J. Hydrogen Energy 2011, 36, 10080–10088. [Google Scholar] [CrossRef]
  10. Fagbenle, R.; Oguaka, A.; Olakoyejo, O. A thermodynamic analysis of a biogas-fired integrated gasification steam injected gas turbine (BIG/STIG) plant. Appl. Therm. Eng. 2007, 27, 2220–2225. [Google Scholar] [CrossRef]
  11. Galvagno, A.; Chiodo, V.; Urbani, F.; Freni, F. Biogas as hydrogen source for fuel cell applications. Int. J. Hydrogen Energy 2013, 38, 3913–3920. [Google Scholar] [CrossRef]
  12. Shan, X.; Qian, Y.; Zhu, L.; Lu, X. Effects of EGR rate and hydrogen/carbon monoxide ratio on combustion and emission characteristics of biogas/diesel dual fuel combustion engine. Fuel 2016, 181, 1050–1057. [Google Scholar] [CrossRef]
  13. Cacua, K.; Amell, A.; Cadavid, F. Effects of oxygen enriched air on the operation and performance of a diesel-biogas dual fuel engine. Biomass Bioenergy 2012, 45, 159–167. [Google Scholar] [CrossRef]
  14. Baratieri, M.; Baggio, P.; Bosio, B.; Grigiante, M.; Longo, G. The use of biomass syngas in IC engines and CCGT plants: A comparative analysis. Appl. Therm. Eng. 2009, 29, 3309–3318. [Google Scholar] [CrossRef]
  15. Fischer, M.; Jiang, X. An investigation of the chemical kinetics of biogas combustion. Fuel 2015, 150, 711–720. [Google Scholar] [CrossRef]
  16. Somehsaraei, H.; Majoumerd, M.; Breuhaus, P.; Assadi, M. Performance analysis of a biogas-fueled micro gas turbine using a validated thermodynamic model. Appl. Therm. Eng. 2014, 66, 181–190. [Google Scholar] [CrossRef]
  17. Hosseini, S.; Bagheri, G.; Wahid, M. Numerical investigation of biogas flameless combustion. Energy Convers. Manag. 2014, 81, 41–50. [Google Scholar] [CrossRef]
  18. Zhen, H.; Leung, C.; Cheung, C. Effects of hydrogen addition on the characteristics of a biogas diffusion flame. Int. J. Hydrogen Energy 2013, 38, 6874–6881. [Google Scholar] [CrossRef]
  19. Wei, Z.; Leung, C.; Cheung, C.; Huang, Z. Effects of equivalence ratio, H2, and CO2 addition on the heat release characteristics of premixed laminar biogas-hydrogen flame. Int. J. Hydrogen Energy 2016, 41, 6567–6580. [Google Scholar] [CrossRef]
  20. Mameri, A.; Tabet, F. Numerical investigation of counter-flow diffusion flame of biogas-hydrogen blends: Effects of biogas composition, hydrogen enrichment and scalar dissipation rate on flame structure and emissions. Int. J. Hydrogen Energy 2016, 41, 2011–2022. [Google Scholar] [CrossRef]
  21. Zhen, H.; Leung, C.; Cheung, C.; Huang, Z. Combustion characteristic and heating performance of stoichiometric biogas-hydrogen-air flame. Int. J. Heat Mass Transf. 2016, 92, 807–814. [Google Scholar] [CrossRef]
  22. Zhen, H.; Leung, C.; Cheung, C.; Huang, Z. Characterization of biogas-hydrogen premixed flames using Bunsen burner. Int. J. Hydrogen Energy 2014, 39, 13292–13299. [Google Scholar] [CrossRef]
  23. Barlow, R. Radiation Models, Radiation, 2003. Available online: http://www.ca.sandia.gov/TNF/radiation.html (accessed on 10 February 2016).
  24. Spalding, D. New Developments and Computed Results, Imperial College CFDU Report HTS/81/1; Imperial College: London, UK, 1980. [Google Scholar]
  25. Patankar, S. Numerical Heat Transfer and Fluid Flow; Hemisphere, CRC press: New York, NY, USA, 1980. [Google Scholar]
  26. Smith, G. GRI-Mech. 1999. Available online: http://combustion.berkeley.edu/gri-mech/ (accessed on 23 November 2015).
  27. Wu, L.; Kobayashi, N.; Li, Z.; Huang, H.; Li, J. Emission and heat transfer characteristics of methane-hydrogen hybrid fuel laminar diffusion flame. Int. J. Hydrogen Energy 2015, 40, 9579–9589. [Google Scholar] [CrossRef]
  28. Tang, C.; Huang, Z.; Law, C. Determination, correlation, and mechanistic interpretation of effects of hydrogen addition on laminar flame speeds of hydrocarbon-air mixtures. Proc. Combust. Inst. 2011, 33, 921–928. [Google Scholar] [CrossRef]
  29. Li, J.; Huang, H.; Kobayashi, N.; Wang, C.; Yuan, H. Numerical study on laminar burning velocity and ignition delay time of ammonia flame with hydrogen addition. Energy 2017, 126, 769–809. [Google Scholar] [CrossRef]
  30. Li, Z.; Han, W.; Liu, D.; Chen, Z. Laminar flame propagation and ignition properties of premixed iso-octane/air with hydrogen. Fuel 2015, 158, 443–450. [Google Scholar] [CrossRef]
  31. Hu, G.; Zhang, S.; Li, Q.; Pan, X.; Liao, S.; Wang, H.; Yang, C.; Wei, S. Experimental investigation on the effects of hydrogen addition on thermal characteristics of methane/air premixed flame. Fuel 2014, 115, 232–240. [Google Scholar] [CrossRef]
Figure 1. Geometry of the combustion chamber (not to scale).
Figure 1. Geometry of the combustion chamber (not to scale).
Energies 10 01200 g001
Figure 2. Comparison of predication temperature (line) and measurement temperature (plot) at the outlet of BG-2/air flames with various H2 addition ratios (black line) and O2 enrichment levels (red line).
Figure 2. Comparison of predication temperature (line) and measurement temperature (plot) at the outlet of BG-2/air flames with various H2 addition ratios (black line) and O2 enrichment levels (red line).
Energies 10 01200 g002
Figure 3. Effect of CH4 content in biogas fuel on combustion temperature distribution.
Figure 3. Effect of CH4 content in biogas fuel on combustion temperature distribution.
Energies 10 01200 g003
Figure 4. Effect of CH4 content in biogas fuel on net heat release rate distribution.
Figure 4. Effect of CH4 content in biogas fuel on net heat release rate distribution.
Energies 10 01200 g004
Figure 5. Effects of H2 addition ratio (a) and O2 enrichment levels (b) in BG-2/air flame on the distribution of combustion temperature.
Figure 5. Effects of H2 addition ratio (a) and O2 enrichment levels (b) in BG-2/air flame on the distribution of combustion temperature.
Energies 10 01200 g005
Figure 6. Effects of H2 addition ratio (a) and O2 enrichment levels (b) in BG-2/air flame on the distribution of the heat release rate.
Figure 6. Effects of H2 addition ratio (a) and O2 enrichment levels (b) in BG-2/air flame on the distribution of the heat release rate.
Energies 10 01200 g006
Figure 7. Heat release rate of biogas flames under H2-enriched (a) and O2-enriched (b) conditions along the centerline.
Figure 7. Heat release rate of biogas flames under H2-enriched (a) and O2-enriched (b) conditions along the centerline.
Energies 10 01200 g007
Figure 8. Effects of H2 addition ratio (a) and O2 enrichment levels (b) in BG-2/air flame on the peak heat release height and average temperature Ta.
Figure 8. Effects of H2 addition ratio (a) and O2 enrichment levels (b) in BG-2/air flame on the peak heat release height and average temperature Ta.
Energies 10 01200 g008
Figure 9. Net reaction rate of biogas flame along centerline at various H2 addition ratios (a) and O2 enrichment levels (b).
Figure 9. Net reaction rate of biogas flame along centerline at various H2 addition ratios (a) and O2 enrichment levels (b).
Energies 10 01200 g009
Figure 10. Comparison of typical elementary reaction exothermic reaction rates of biogas flame along the centerline at H2 addition ratio of 0% (a) and 50% (b).
Figure 10. Comparison of typical elementary reaction exothermic reaction rates of biogas flame along the centerline at H2 addition ratio of 0% (a) and 50% (b).
Energies 10 01200 g010
Figure 11. Flame thickness along the centerline at various H2 addition ratios (black line) and O2 enrichment levels (red line).
Figure 11. Flame thickness along the centerline at various H2 addition ratios (black line) and O2 enrichment levels (red line).
Energies 10 01200 g011
Table 1. Composition of biogas flames at H2-enriched and O2-enriched conditions.
Table 1. Composition of biogas flames at H2-enriched and O2-enriched conditions.
ItemsCH4/L·min−1CO2/L·min−1H2/L·min−1O2/L·min−1Air/L·min−1Power/kW
BG-11.8291.220--18.3401
BG-21.8290.784--18.3401
BG-31.8290.457--18.3401
H2 10%1.6471.0980.602-18.0141
H2 30%1.2810.8541.805-17.3611
H2 50%0.9150.6103.008-16.7091
O2 25%1.8291.220-0.78014.6261
O2 30%1.8291.220-1.46311.3761
O2 35%1.8291.220-1.9509.0541
Table 2. Absorption coefficient of a0 to a5 for CH4, CO, CO2, and H2O.
Table 2. Absorption coefficient of a0 to a5 for CH4, CO, CO2, and H2O.
CoefficientCO2H2OCH4CO (T < 750 K)CO (T > 750 K)
a018.741−0.230936.63344.786910.09
a1−121.31−1.1239−3.5686 × 10−3−0.06953−0.01183
a2273.59.41531.6682 × 10−82.95775 × 10−44.7753 × 10−6
a3−194.05−2.99882.5611 × 10−10−4.25732 × 10−7−5.87209 × 10−10
a456.310.51382−2.6558 × 10−142.02894× 10−10−2.5334 × 10−14
a5−5.8169−1.884 × 10−50--

Share and Cite

MDPI and ACS Style

Li, J.; Huang, H.; Huhetaoli; Osaka, Y.; Bai, Y.; Kobayashi, N.; Chen, Y. Combustion and Heat Release Characteristics of Biogas under Hydrogen- and Oxygen-Enriched Condition. Energies 2017, 10, 1200. https://doi.org/10.3390/en10081200

AMA Style

Li J, Huang H, Huhetaoli, Osaka Y, Bai Y, Kobayashi N, Chen Y. Combustion and Heat Release Characteristics of Biogas under Hydrogen- and Oxygen-Enriched Condition. Energies. 2017; 10(8):1200. https://doi.org/10.3390/en10081200

Chicago/Turabian Style

Li, Jun, Hongyu Huang, Huhetaoli, Yugo Osaka, Yu Bai, Noriyuki Kobayashi, and Yong Chen. 2017. "Combustion and Heat Release Characteristics of Biogas under Hydrogen- and Oxygen-Enriched Condition" Energies 10, no. 8: 1200. https://doi.org/10.3390/en10081200

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