progress in energy and combustion science · 2019-12-31 · post-combustion capture:co2 is...

59
Review Oxy-fuel combustion of pulverized coal: Characterization, fundamentals, stabilization and CFD modeling Lei Chen, Sze Zheng Yong, Ahmed F. Ghoniem * Department of Mechanical Engineering, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, MA 02139-4307, USA article info Article history: Received 13 January 2011 Accepted 27 August 2011 Available online 14 December 2011 Keywords: Carbon capture Oxy-fuel combustion Coal Heat transfer Flame stabilization CFD abstract Oxy-fuel combustion has generated signicant interest since it was proposed as a carbon capture technology for newly built and retrotted coal-red power plants. Research, development and demonstration of oxy- fuel combustion technologies has been advancing in recent years; however, there are still fundamental issues and technological challenges that must be addressed before this technology can reach its full potential, especially in the areas of combustion in oxygen-carbon dioxide environments and potentially at elevated pressures. This paper presents a technical review of oxy-coal combustion covering the most recent experi- mental and simulation studies, and numerical models for sub-processes are also used to examine the differences between combustion in an oxidizing stream diluted by nitrogen and carbon dioxide. The evolution of this technology from its original inception for high temperature processes to its current form for carbon capture is introduced, followed by a discussion of various oxy-fuel systems proposed for carbon capture. Of all these oxy-fuel systems, recent research has primarily focused on atmospheric air-like oxy-fuel combustion in a CO 2 -rich environment. Distinct heat and mass transfer, as well as reaction kinetics, have been reported in this environment because of the difference between the physical and chemical properties of CO 2 and N 2 , which in turn changes the ame characteristics. By tracing the physical and chemical processes that coal particles experience during combustion, the characteristics of oxy-fuel combustion are reviewed in the context of heat and mass transfer, fuel delivery and injection, coal particle heating and moisture evap- oration, devolatilization and ignition, char oxidation and gasication, as well as pollutants formation. Operation under elevated pressures has also been proposed for oxy-coal combustion systems in order to improve the overall energy efciency. The potential impact of elevated pressures on oxy-fuel combustion is discussed when applicable. Narrower ammable regimes and lower laminar burning velocity under oxy-fuel combustion conditions may lead to new stability challenges in operating oxy-coal burners. Recent research on stabilization of oxy-fuel combustion is reviewed, and some guiding principles for retrot are summarized. Distinct characteristics in oxy-coal combustion necessitate modications of CFD sub-models because the approximations and assumptions for air-fuel combustion may no longer be valid. Advances in sub-models for turbulent ow, heat transfer and reactions in oxy-coal combustion simulations, and the results obtained using CFD are reviewed. Based on the review, research needs in this combustion technology are suggested. Ó 2011 Elsevier Ltd. All rights reserved. Contents 1. Introduction ....................................................................................................................... 159 1.1. Carbon capture technologies for coal-fired power plants ....................................... .................................. 159 1.2. Focus and methods .......................................................................................................... 161 1.3. Brief history of oxy-fuel combustion technologies .......................................... ..................................... 161 1.3.1. Oxy-fuel combustion for high temperature processes ..................................................................... 161 1.3.2. Oxy-fuel combustion for CCS ........................................................................................... 162 1.4. Oxy-coal combustion systems ................................................................................................. 163 1.4.1. Atmospheric oxy-coal combustion systems with flue gas recycle ........................................................... 163 1.4.1.1. Air Separation Unit (ASU) .................................................................................... 163 * Corresponding author. Tel.: þ1 617 253 2295; fax: þ1 617 253 5981. E-mail address: [email protected] (A.F. Ghoniem). Contents lists available at SciVerse ScienceDirect Progress in Energy and Combustion Science journal homepage: www.elsevier.com/locate/pecs 0360-1285/$ e see front matter Ó 2011 Elsevier Ltd. All rights reserved. doi:10.1016/j.pecs.2011.09.003 Progress in Energy and Combustion Science 38 (2012) 156e214

Upload: others

Post on 31-May-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

at SciVerse ScienceDirect

Progress in Energy and Combustion Science 38 (2012) 156e214

Contents lists available

Progress in Energy and Combustion Science

journal homepage: www.elsevier .com/locate/pecs

Review

Oxy-fuel combustion of pulverized coal: Characterization, fundamentals,stabilization and CFD modeling

Lei Chen, Sze Zheng Yong, Ahmed F. Ghoniem*

Department of Mechanical Engineering, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, MA 02139-4307, USA

a r t i c l e i n f o

Article history:Received 13 January 2011Accepted 27 August 2011Available online 14 December 2011

Keywords:Carbon captureOxy-fuel combustionCoalHeat transferFlame stabilizationCFD

* Corresponding author. Tel.: þ1 617 253 2295; faxE-mail address: [email protected] (A.F. Ghoniem)

0360-1285/$ e see front matter � 2011 Elsevier Ltd.doi:10.1016/j.pecs.2011.09.003

a b s t r a c t

Oxy-fuel combustion has generated significant interest since itwasproposed as a carbon capture technologyfor newly built and retrofitted coal-fired power plants. Research, development and demonstration of oxy-fuel combustion technologies has been advancing in recent years; however, there are still fundamentalissues and technological challenges thatmustbe addressedbefore this technologycanreach its full potential,especially in the areas of combustion in oxygen-carbon dioxide environments and potentially at elevatedpressures. This paper presents a technical review of oxy-coal combustion covering the most recent experi-mental and simulation studies, and numerical models for sub-processes are also used to examine thedifferences between combustion in an oxidizing stream diluted by nitrogen and carbon dioxide. Theevolution of this technology from its original inception forhigh temperature processes to its current form forcarbon capture is introduced, followed by a discussion of various oxy-fuel systems proposed for carboncapture. Of all these oxy-fuel systems, recent research hasprimarily focused on atmospheric air-like oxy-fuelcombustion in a CO2-rich environment. Distinct heat and mass transfer, as well as reaction kinetics, havebeen reported in this environment becauseof thedifferencebetween thephysical andchemical propertiesofCO2 and N2, which in turn changes the flame characteristics. By tracing the physical and chemical processesthat coal particles experience during combustion, the characteristics of oxy-fuel combustion are reviewed inthe context of heat and mass transfer, fuel delivery and injection, coal particle heating and moisture evap-oration, devolatilization and ignition, char oxidation and gasification, as well as pollutants formation.Operation under elevated pressures has also been proposed for oxy-coal combustion systems in order toimprove the overall energy efficiency. The potential impact of elevated pressures on oxy-fuel combustion isdiscussedwhenapplicable.Narrowerflammable regimesand lower laminarburningvelocityunderoxy-fuelcombustion conditions may lead to new stability challenges in operating oxy-coal burners. Recent researchonstabilizationof oxy-fuel combustion is reviewed, andsomeguidingprinciples for retrofit are summarized.Distinct characteristics in oxy-coal combustion necessitate modifications of CFD sub-models because theapproximations andassumptions forair-fuel combustionmayno longerbevalid. Advances in sub-models forturbulent flow, heat transfer and reactions in oxy-coal combustion simulations, and the results obtainedusing CFD are reviewed. Based on the review, research needs in this combustion technology are suggested.

� 2011 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1591.1. Carbon capture technologies for coal-fired power plants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1591.2. Focus and methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1611.3. Brief history of oxy-fuel combustion technologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161

1.3.1. Oxy-fuel combustion for high temperature processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1611.3.2. Oxy-fuel combustion for CCS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162

1.4. Oxy-coal combustion systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1631.4.1. Atmospheric oxy-coal combustion systems with flue gas recycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163

1.4.1.1. Air Separation Unit (ASU) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163

: þ1 617 253 5981..

All rights reserved.

Page 2: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 157

1.4.1.2. Carbon dioxide purification unit (CPU) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1641.4.1.3. Flue gas recycle (FGR) system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164

1.4.2. Pressurized oxy-coal combustion systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1641.4.3. Performance of the oxy-coal combustion systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165

2. Thermodynamics, transport and chemistry in oxy-coal combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1662.1. Heat transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166

2.1.1. Radiative heat transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1662.1.2. Convective heat transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1672.1.3. Matching of combustion temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1682.1.4. Matching of radiative and convective heat transfer distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171

2.2. Fuel delivery and injection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1722.2.1. Pulverized coal transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1722.2.2. Coal-water slurry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172

2.3. Heating and moisture evaporation of coal particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1732.4. Coal devolatilization and char formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1752.5. Ignition of coal particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176

2.5.1. Theoretical analysis of homogenous ignition delay . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1762.5.2. Experimental studies on coal particle ignition delay in O2/N2 and O2/CO2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1772.5.3. Burning of volatiles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179

2.6. Oxy-char combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1792.6.1. Kinetics of char oxidation in oxy-fuel combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1802.6.2. Effect of diffusivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1812.6.3. Effect of gasification reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1832.6.4. Effect of heat capacity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1852.6.5. Effect of pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1852.6.6. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186

2.7. Pollutants formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1872.7.1. Fly ash formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1872.7.2. Ash deposit/slag formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1882.7.3. NOx formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188

2.7.3.1. Nitrogen chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1882.7.3.2. Lower NOx emission intensities in oxy-coal combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1892.7.3.3. Reasons for lower NOx emission intensities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190

2.7.4. SO2 formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1912.7.5. CO concentration and emission . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191

2.8. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1923. Stabilization of oxy-coal combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .193

3.1. Flammability under oxy-fuel conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1933.2. Laminar flame propagation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193

3.2.1. Effect of physical properties and chemical kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1943.2.2. Impact of operating conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196

3.3. Stabilizing oxy-coal combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1973.3.1. Optimized flue gas recycle ratios . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1973.3.2. Oxygen feeding strategies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197

3.3.2.1. Pure oxygen injection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1973.3.2.2. Oxygen partitioning in different streams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199

3.3.3. Fluid dynamics considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2013.3.3.1. Matched mass flow rate, momentum or velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2013.3.3.2. Advanced burner design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202

3.3.4. Gas stream preheating . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2024. CFD modeling of oxy-coal combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202

4.1. Governing equations and physical properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2024.2. Turbulent flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2044.3. Radiation heat transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205

4.3.1. Models for radiative properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2054.3.2. Modification of the gray-gas model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205

4.4. Heterogeneous reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2064.5. Gas phase reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207

4.5.1. Reduced reaction mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2074.5.2. Turbulence-chemistry interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207

4.6. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2085. Research need . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208

5.1. Oxy-coal system design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2085.2. Fundamentals of oxy-coal combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2085.3. CFD modeling of oxy-coal combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2095.4. Large scale demonstration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209Acknowledgment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210

Page 3: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Nomenclature

SymbolsA pre-exponential factor for reaction kineticsA area (m2)a averaged absorption coefficient (1=m)C concentration (kmol=m3)ci mass diffusion-limited constant (s=K0:75)cp specific heat at constant pressure (J=kgK)cv molar heat capacity at constant volume (J/kmolK)rCp volumetric heat capacity at constant volume (J=m3K)D diameter of pipe (m)DAB mass diffusivity (m2=s)d diameter of droplet or particle (m)E activation energy (kJ=mol)g gravity (9:81 m=s2)h convective heat transfer coefficient (W=m2K)hfg latent heat (J=kg)hm convective mass transfer coefficient (m=s)Dh reaction heat of char oxidation (J=kg)I radiation intensity (W=m2 � sr)k thermal conductivity (W=mK)k reaction kinetics kg=ðm2$Pa$sÞk turbulence kinetic energyMw;MW molecular weight (g=mol)_m mass consumption rate (kg=s)Nu Nusselt number_n mole flow rate (mol=s)P pressure (Pa)Pr Prandtl numberq heat transfer (W)q00 heat flux (W=m2)qc heat release by combustion reaction (J=kmol)Ru ideal gas constant (8314J=ðkmol,KÞ)Re Reynolds numberr radius (m)s beam length in radiation (m)S source termSc Schmidt numberSh Sherwood numberSL flame propagation velocity (m=s)T temperature (K)u velocity (m=s)V volume (m3)v velocity (m=s)We Weber numberY mass fraction

Greek lettersa thermal diffusivity (m2=s)a absorptivityε emissivityε gas fraction contributing to atomizationε dissipation ratek absorption coefficient (1=m)k pressure absorption coefficient (1=m� atm)l wavelength of radiation (mm)n kinematic viscosity (m2=s)n mass ratio of oxygen-to-fueln stoichiometric coefficient of carbon in heterogeneous

reactionsr density (kg=m3)

s interfacial surface tension (N=m)s StefaneBoltzmann constant (5:67� 10�8 W=m2K4)ss scattering coefficients time (s)F Favre-averaged variableF phase function in RTEf equivalence ratioU solid angle

Subscripts0 initial value at time 0, reference valuea activationair air-fuel combustion, or O2/N2

blowout blowout velocityboil boiling stateC charc combustionconv convectiond diameter of particle or dropletd devolatilizationdiff diffusionF fuelf flameg gasI inert gasi ignitioni gas component ikin kineticsl liquidmix gas mixtureoxy oxy-fuel combustion, or O2/CO2

P productp particlepg particle-gaspu pickup velocitypw particle-wallR reactantrad radiations surfacesur surrounding gasu unburned gasw wallN infinity

AcronymsAFT adiabatic flame temperatureANL Argonne national laboratoryASU air separation unitB&W Babcock & Wilcox companyCANMET Canada Center for Mineral and Energy TechnologyCCS carbon capture and sequestrationCPU carbon dioxide purification unitCWS coal water slurryCFD computational fluid dynamicsDAEM distributed activation energy modelDFM/SFM double/single film modelDNS direct numerical simulationDTF drop tube furnaceDR dilution ratioEFR entrained flow reactorEOR enhanced oil recoveryEWB exponential wide bandESP electrostatic precipitator

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214158

Page 4: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

FGD flue gas desulphurizationGLR gas-to-liquid ratioHRSG heat recovery steam generatorHiTAC high temperature air combustionHTC heat transfer coefficientITM ion transport membranesLES large eddy simulationLHV lower heating valueMILD moderate or intense low-oxygen dilutionNETL national energy technology laboratoryOEC oxygen enriched/enhanced combustionPC pulverized coalPFG product flue gas

RANS Reynolds-Averaged Navier-StokesRFG recycled flue gasRR recycle ratioSCR selective catalytic reductionSMD Sauter mean diameterSNBM statistical narrow band modelTGA thermogravimetric analysisUBC unburned carbonWMR wire mesh reactorWSGG weighted-sum-of-gray-gasesWSR well stirred reactorZEPP zero emission power plants

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 159

1. Introduction

1.1. Carbon capture technologies for coal-fired power plants

Reliable, efficient and clean energy supply is one of the basicneeds of humankind. Today, our energy supply system is under-going a long-term transition from its conventional form to a moresustainable and low carbon style, especially addressing greenhousegas (water, carbon dioxide, methane, nitrous oxide, chlorofluoro-carbons and aerosols) emissions into the atmosphere. Strongevidence suggests that both the average global temperature and theatmospheric CO2 concentration have significantly increased sincethe onset of the industrial evolution, and they are well correlated[1]. Concerns over climate change have led to mounting efforts ondeveloping technologies to reduce carbon dioxide emissions fromhuman activities [2,3]. Technological solutions to this problemought to include a substantial improvement in energy conversionand utilization efficiencies, carbon capture and sequestration (CCS),and expanding the use of nuclear energy and renewable sourcessuch as biomass, hydro-, solar, wind and geothermal energy [2].

Coal has been and will continue to be one of the major energyresources in the long term because of its abundant reserves andcompetitively low prices, especially for the use of base-load powergeneration. For instance, the share of coal in world energyconsumption was 29.4% in 2009, as opposed to 34.8% for oil and23.8% for natural gas [4]. In terms of power generation, coalcontinues to be the dominant fuel, contributing about 45% of thetotal electricity in the US in 2009 [5], and about 80% in China.Several technologies have been proposed for reducing CO2 emis-sion from coal-fired power generation, namely post-combustioncapture, pre-combustion capture and oxy-fuel combustioncapture [6]:

� Pre-combustion capture: Fuel is either gasified or reformed tosyngas, a mixture of carbon monoxide and hydrogen, which isthen shifted via steam reforming. CO2 is then separated fromthe syngas by shifting carbon monoxide with steam, yieldingpure hydrogen (water gas shift reaction). The Integrated Gasi-fication Combined Cycles (IGCC) for coal is an example of pre-combustion capture system.

� Post-combustion capture: CO2 is separated from the flue gasesusing chemical solvents [7], sorbents (such as calcium oxide [8]or carbon fibers [9]) andmembranes [10] without changing thecombustion process. However, the addition of a post-combustion capture unit may change the steam cycle becauselarge quantity of low pressure steam must be extracted fromthe steam cycle for the solvent regeneration process.

� Oxy-fuel combustion: Instead of using air as oxidizer, pureoxygen (O2) or a mixture of O2 and recycled flue gas is used togenerate high CO2 concentration product gas; therefore, thecombustion process is significantly changed. Chemical-LoopingCombustion (CLC) is another combustion process that belongsto the oxy-fuel combustion category, in which pure oxygenrather than air is supplied bymetal oxides for combustion, suchthat the mixing between CO2 and N2 is inherently avoided. Thistechnology is not the primary focus of this paper, and, thereader is referred to [11e13] for more details on CLC.

In general, the technologies described above can be applied togenerate energy from natural gas and coal with the exemption ofsome low rank coals due to unresolved engineering challenges,however, because of the important role of pulverized coal in baseload electricity generation and its contribution to CO2 emission, thispaper is primarily concerned with the combustion of pulverizedcoal, although some mention is made of other fuels as well.

These three major carbon capture technologies for coal-firedpower plants have been studied in terms of power generationefficiency, capital costs and costs of electricity (COE) [14e16].Representative energy efficiency and economic performance ofthese technology options are compared in Table 1. All of theseestimates are based on 90% CO2 capture in rebuilt and retrofittedscenarios. The cost of CO2 indicates the cost that is incurred tocapture 1 metric ton carbon dioxide without transportation andstorage. Although the absolute numbers vary by few percentagepoints in these studies, all reports show the same trends. In general,all three capture technologies result in an efficiency penalty, whileoxy-fuel capture and pre- capture or IGCC show advantages overpost-combustion capture in terms of COE and cost of CO2. The IGCCtechnology yields a higher generation efficiency and a slightlylower cost than oxy-fuel combustion technology. However, allthese technologies are in their early stages of development and stillhave great potential for improvement.

In particular, these studies have a common conclusion that oxy-fuel combustion is the most competitive technology option forretrofitting existing coal-fired power plants, which at the momenthave the largest potential for CCS. Fig. 1(a) shows the installedpower capacity of existing coal-fired power plants and their initialoperation year in the U.S. Although the number of newly-built coalpower generation units declined since 1990s’, there is a resurgenceof new coal power plants in recent years. Moreover, about 98.7 GWor 29% of all the existing coal-fired power capacity were built after1980 [17]. This situation is even more prominent in developingcountries such as China and India, where the coal power generationcapacity has been booming in the last two decades. As shown in

Page 5: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Table 1Representative performance and economics data for the three main capture technologies, from [14].

Performance Supercritical PCa SCb PC-oxyfuel IGCCc

w/o capture w/capture w/capture w/o capture W/capture

Generating efficiency 38.5% 29.3% 30.6% 38.4% 31.2%Efficiency penalty CO2 recovery (heat): �5% Boiler/FGD: 3% Water/Gas shift: �4.2%

CO2 compression: �3.5% ASU: �6.4% CO2 compression: �2.1%CO2 recovery (power): �0.7% CO2 compression: �3.5% CO2 recovery: �0.9%

Other: �1%Capital Cost ($/kWe)e 1330 2140 (1314)d 1900 (867)d 1430 1890COE (c/kWh)e 4.78 7.69 6.98 5.13 6.52Cost of CO2 ($/t)e 40.4 30.3 24.0

a PC: pulverized coal.b SC: supercritical.c IGCC: Integrated gasification combined cycle.d Figures in parenthesis are the expected capital cost for retrofits.e Based on design studies done between 2000 & 2004, a period of cost stability, updated to 2005$ using CPI inflation rate.

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214160

Fig. 1(b), the built rate of coal-fired power generation units in Chinais up to 80 GW per year, significantly higher than that of the U.S.This growth is driven by China’s fast economic development [18]. Itcan safely be assumed that a sizable reduction of CO2 emission fromexisting plants would come from retrofits. Oxy-fuel combustionsystems have a natural advantage in retrofitting existing PC powerplants because they can reusemost of the existing plant equipment.

Fig. 1. Coal-fired power capacity as function of the initial year of operation in U.S. and China:those after 1960s). Data are cited from EIA-860 annual electric generator report [17] (b) NewChina, figure courtesy NETL study [18].

The advantages of oxy-fuel combustion as a retrofit technology arealso indicated in Table 1. The capital cost for supercritical PCretrofits with oxy-fuel is $867/kWe, which is significantly lowerthan the capital cost of post-combustion retrofit ($1314/kWe) andof newly-built IGCC plants ($1890/kWe).

Considering the advantages of a relatively moderate efficiencypenalty and the lowest retrofit capital expenditure, atmospheric

(a) Existing coal-fired power capacity by initial year of operation in the U.S. (only showsly-built and planned coal-fired power generation capacity since year 2000 in US. and

Page 6: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Fig. 2. Schematic diagram of oxy-fuel combustion regimes as a function of the oxygenmole fraction and the preheat temperature of the reactant stream.

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 161

oxy-fuel combustion systems have been widely accepted asa competitive carbon capture technology. More recently, it has beenadopted to substitute the original IGCC plan in the U.S. DOEFutureGen 2.0 program [19]. Previous studies have reviewed itscharacteristics [6,20e22] as well as recent developments in pilot-scale and commercial-scale demonstration plants [23]. Whilesuccessful, the technology still faces many challenges, such as airleakage into the flue gas system, the relatively low energy effi-ciency, the need for efficient air separation and better plant inte-gration and flue gas cleanup, among others. In particular, significantchallenges are expected in the combustion process itself, includingstability and emissions, burner design and scaling, as well asdetermining of optimal operating conditions. This paper focuses onthe combustion process, especially the impact of the absence ofnitrogen and the reuse of flue gas as a diluent on the physics,chemistry and dynamics of oxy-combustion.

1.2. Focus and methods

The paper focuses on reviewing the fundamentals of oxy-fuelcombustion of pulverized coal in a CO2-rich environment. Inparticular, it examines how oxy-fuel combustion differs from air-fuel combustion in areas including fuel delivery, slurry atomiza-tion, heat transfer, coal particle combustion and its interactionswith the gaseous phase, flame stabilization and pollutant forma-tion. Aside from reviewing experimental and numerical studies inthe literature, qualitative modeling studies using simplified modelsare also carried out with the objective of determining the impact ofthe dilution medium on the combustion process. In these modelingstudies, detailed chemical reaction kinetics and transport proper-ties in a well stirred reactor or a plug flow reactor, as well assimplified numerical models for single coal particle combustion,are used. The implications of the simulation results on combustordesign are discussed. Another focus of this review is the applica-tions of oxy-coal combustion at elevated pressures. The potentialimpact of the pressure on the oxy-combustion characteristics isdiscussed based on previous relevant studies. These discussionswill enable future developments of oxy-coal combustion overa wider range of operating conditions.

The paper outline is as follows: following an introduction on theevolution and characterization of oxy-fuel combustion and itsapplications in power plants for CCS in the introduction Section 1,in Section 2, some of the fundamental processes in oxy-coalcombustion are discussed. The contents are organized in sucha way that each subsection starts with a physical description of theprocess or a scaling analysis showing the difference between oxy-coal and air-coal combustion. Experimental and numericalstudies on oxy-combustion of coal and at times, of gaseous fuels,are reviewed in order to compare the combustion phenomena inO2/CO2 and in O2/N2 environments. These comparisons include theimpact of the environment on heat transfer, both locally andglobally, transportation, demoisturization, devolatilization andignition of the coal particles, char burning, as well as combustion inthe homogeneous phase, etc. Research on pollutant formationunder oxy-fuel conditions is also briefly reviewed. However, thereader is referred to other literature reviews in more detail.

The distinct characteristics of oxy-coal combustion raise newchallenges in flame stabilization, and this issue is addressed inSection 3. Using simplified models, the flammability limit andlaminar burning velocity under oxy-fuel combustion conditions areshown to be different than in air-fuel combustion, indicating thatdestabilization problems may occur when using conventional coalburners. The impact of the fuel-oxygen equivalence ratio, flue gasrecirculation and preheat temperature on flammability areanalyzed. Based on a review of experimental studies on oxy-coal

flame stabilization, the general principles for oxy-coal burnersretrofitting and designing are discussed. Many of the local, meso-scale and large-scale models discussed earlier in the paper are usedin computational fluid dynamics (CFD) simulations of coalcombustion. The modification of sub-models and modeling resultsof numerical studies are depicted in Section 4. Finally, someresearch needs in the field are proposed in Section 5.

1.3. Brief history of oxy-fuel combustion technologies

1.3.1. Oxy-fuel combustion for high temperature processesEven before the concern over CO2 emissions arose, forms of oxy-

fuel combustion had been used in different applications. Fig. 2shows schematically oxy-fuel combustion regimes as function ofthe preheat temperature and oxygen mole fraction of the oxidantstream. The following regimes have been identified:

� Zone I: Air combustion (diluted with N2), or air-like oxy-fuelcombustionwith a flue gas recycle ratio (RR) of 60e80% vol (i.e.oxygen mole fraction of about 30% in the oxidizer stream). Therecycle ratio is defined as:

RR ¼ mRFG

mRFG þmPFG(1)

where mRFG is the recycled gas mass flow rate and mPFG is theproduct gas mass flow rate respectively.

� Zone II: Oxygen-enriched or oxygen-enhanced combustion(OEC) with oxygen mole fraction significantly higher than 21%.

� Zone III: Full oxy-fuel combustion with pure oxygen.� Zone IV: High temperature air combustion (HiTAC) and hightemperature oxy-fuel combustion with greatly preheatedoxidizer.

In existing utility boilers, pulverized coal is typically combustedin air which consists of 21% O2 and 79% N2, placing this combustionregime in Zone I. However, the flame temperature of air combus-tion is not high enough for some industrial processes such as theproduction of glass, metal and cement. Therefore, oxy-enhanced oroxy-enriched combustion (Zone II), featuring a higher flametemperature and a higher available sensible enthalpy, wasproposed in the 1970s to improve the energy efficiency in these

Page 7: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214162

processes [24]. At that time, oxy-fuel combustion was defined ascombustion with an oxidizer stream containing oxygen mole frac-tions higher than 21%, in which nitrogen was retained as the dilu-tion gas. Because of its advantages in NOx reduction and costeffectiveness in some applications, this technologywas first appliedin the glass and metal industries and then in the cement andincineration industries as an alternative to the High TemperatureAir Combustion (HiTAC) technology (Zone IV) [25,26].

As the combustion regime moves from zone II toward the rightin Fig. 3, i.e., as the oxygen mole fraction is increased, thecombustion intensity rises. As the dilution is further decreased tozero, the regime achieved is called the full oxy-fuel combustionregime (Zone III). In this regime, pure oxygen is used, resulting ina very intense flame. An example of a typical full oxy-fuelcombustion system is the acetylene oxy-fuel flame used for weld-ing and cutting purposes. The adiabatic flame temperature is ashigh as 3343 K, which is much higher than that of the air-combustion (2535 K) due to the absence of inert gases that dilutethe products in air-combustion [24].

The combustion regime at the top of Fig. 2 is the high temper-ature air combustion (HiTAC) regime (Zone IV). High temperaturesare achieved because of the use of highly preheated air. The oxygenmole fraction can be reduced significantly below 21% while stillmaintaining stable combustion. For instance, experimental resultsshow stable liquefied petroleum gas (LPG) flame combustion withlow oxygen concentration air (less than 5%) at air temperaturesabove 800 �C [26]. In recent years, a new combustion technologynamed MILD combustion or flameless combustion has beendeveloped in this regime, featuring a uniform temperature distri-bution and low NOx emissions. This phenomenon is defined byCavaliere and Joannon [27] as “a combustion process with the inlettemperature of the reactant mixture that is higher than mixtureself-ignition temperature, whereas the maximum allowabletemperature increase with respect to inlet temperature duringcombustion is lower than mixture self-ignition temperature”. Sinceno intense flame front is formed in MILD combustion, NOx emis-sions are significantly reduced.

Note that all the “historical” oxy-fuel and HiTAC systems dis-cussed above are developed for high temperature processes and arenot the same as the oxy-fuel combustion systems designed for CCS.Unlike the oxy-fuel combustion for CCS, nitrogen is still used as thediluent gas, therefore, the oxy-fuel combustion used for hightemperature applications are different from oxy-fuel combustion forCCS, and we will limit our discussion to the characteristics of thelatter.

1980 1990 2000 2010 20200.1

1

10

100

1000

Vattenfall 250

Youngdong 100Jamestown 50

Callide A 30

Vattenfall 10

TOTAL (NG) 10

Pearl Plant 22Oxy-coal UK 13.3

CIUDEN 10CIUDEN 6.7

Jupiter 6.7

B&W 10International Comb 11.7

JSIM/NEDO (Oil) 4.0

ENEL 1.0

RWE-NPOWER 0.2

PowerGen 0.3IVD-Stuttgart 0.2CANMET 0.1

B&W/AL 0.4

IHI 0.5

Demonstration with CCSIndustrial scale without CCS

Pilot scale

MW

e (o

r MW

t/3)

Year

ANL/BHP 0.2

ANL/EERC 1.0 IFRF 1.0

Fig. 3. Historical progress of the scale of oxy-fuel pilot plants and demonstrations.Adopted from [23]. The symbol - indicates demonstration plants with CCS, : indi-cates industrial scale tests without CCS, and B indicates pilot scale tests.

1.3.2. Oxy-fuel combustion for CCSWhen nitrogen is removed from the oxidant, a high CO2

concentration stream is generated at the end of the combustionprocess. Usually, recycled flue gases are used to replace nitrogenand control the combustion temperature. The idea of applying oxy-fuel processes with flue gas recycle in coal-fired plants to controlthe CO2 emission [28,29] and/or produce high concentration CO2for enhanced oil recovery (EOR) was first proposed in 1982 [28,30].As such, the main purpose of oxy-fuel combustion applications incoal-fired plants is no longer high temperatures, but the recovery ofconcentrated carbon dioxide in an air-like combustion environ-ment. Following these proposals, Argonne National Laboratory(ANL) pioneered the investigation of this process in the mid andlate 1980s, focusing on the system and its combustion character-istics [31e33]. Soon after, more and more researchers agreed thatthis system complements the two other major approaches forcarbon dioxide capture, which led to a renewed interest in thistechnology in the 1990s. Research conducted by the InternationalFlame Research Foundation (IFRF), CANMET, IHI, as well as otherinstitutes and industrial parties has made considerable contribu-tions in understanding of this process.

In the case of retrofitting existing PC power plants, the combus-tion temperature and heat transfer rate in the conventional aircombustion system are intentionally maintained in order to utilizethe existing equipment. Therefore, the operating conditions of theoptimized oxy-fuel are such that an air-combustion-like environ-ment is achievedwhile satisfying the economic and safety concerns.The air-like oxy-fuel combustion (categorized in Zone I of Fig. 2)system thus has been viewed not only as an appropriate technologyfor new units but also as an excellent retrofit strategy for existingcoal-firedpowerplants. Similar, orhigher,O2 concentrations as in aircombustion systems (15e30% by volume) are used in this regime,and theoxidizer stream is not preheated. The associated combustionphenomena in this regime have been extensively studied andreviewed by Buhre et al. [21], Wall et al. [22] and most recently, byToftegaard et al. [20] and Zheng et al. [34].

Along with the research and development on the air-like oxy-coal technology, pilot and large scale demonstration plants arebeing built around the world. Wall et al. [23] surveyed research onoxy-fuel technology, from pilot-scale tests, to industry-scale testsand full-scale demonstrations, and compiled the historical devel-opment of this technology worldwide, as shown in Fig. 3. The year2008 marks an important milestone with the commissioning of theworld’s first 30 MWth demonstration plant in Germany. Morelarge-scale demonstrations in industry-scale coal-fired boilers havebeen planned or are already underway, as shown in Table 2, basedon the work of Wall et al. [23] and Herzog [35]. Success in thesedemonstrations is expected to lead to wider commercialdeployment.

Recent research has also focused on extending the range ofoperating conditions of oxy-coal combustion to improve energyefficiency, environmental performance and economics of thistechnology. “Full” oxy-fuel combustion was proposed by CANMETfor their third generation oxy-fuel combustion system [36] with thematerial constraints being the main concern and challenge.Although the MILD regime is difficult to realize in coal combustionbecause the heat recuperator used for external heat recirculationcannot handle flue gases with fly ash, efforts have been made toapply this regime to air-fired coal combustion [37e41] and morerecently, to oxy-fired coal combustion [40,42,43]. MILD oxy-coalcombustion regime may be achieved by using highly preheated[37], moderately preheated [42,44] or even oxidizer streamswithout preheating [40]. Internal flue gas recirculation can makeMILD combustion conditions locally possible in the vicinity of theburner [45]. These proposed operating conditions for oxy-coal

Page 8: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Table 2List of ongoing and proposed large scale oxy-coal combustion demonstration projects, modified fromWall et al. [23], the CCS project database of MIT Energy Initiative [35] withcurrent updates of these projects.

Project name Leader Location Scale Technology MWe New/Retrofit Power Gen CO2 Seq Start-up

Jupiter Pearl Plant Jupiter USA Pilot 22(MWth) R N NAc 2007B&W pilot plant B&W USA Pilot PCa 30(MWth) R N N 2008OxyCoal-UK Doosan Babcock UK Pilot PCa 40(MWth) R N N 2009Alstom Windsor Facility Alstom USA Pilot PCa 15(MWth) R N N 2009Schwarze Pumpe Vattenfall Germany Pilot PCa 10 N N Seqd 2008Callide-A CS Energy, IHI etc. Australia Pilot PCa 30 R Y Seqd 2011Compostilla (OXY-CFB-300)Phase I

ENDESA, CIUDEN andFoster Wheeler

Spain Pilot CFBb 17 N Y Seq d 2011e2012

Jamestown Jamestown BPU USA Demo CFB 43 N N Seqd 2013Janschwalde Vattenfall Germany Demo PCa 250 N Y Seq d 2015FutureGen FutureGen Alliance USA Demo PCa 200 R Y Seq d 2015Compostilla (OXY-CFB-300)Phase II

ENDESA, CIUDEN andFoster Wheeler

Spain Demo CFBb 300 N Y Seq d 2015

Youngdong KEPCO S. Korea Demo PCa 100 R Y Seq d 2016Black Hills Power Black Hills Corporation USA Demo PCa 100 N Y NAc 2016

a PC: Pulverized Coal.b CFB: Circulated Fluid Bedc NA: Data Not Availabled Seq: Sequestration

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 163

combustion, however, will require major changes to the heatexchanger system, and are still at their conceptual phases.

1.4. Oxy-coal combustion systems

Theconceptofoxy-coal combustion involves theburningofcoal inpure oxygen to obtain a carbon dioxide-rich stream that is ready forsequestration, after removing water vapor and other impurities.Various oxy-coal combustion systems have been proposed in theliterature [21,42,46e48]. Thefirst version is the atmospheric pressureoxy-coal systeminwhichfluegases arepartially recycled tomoderatethe flame temperatures. An alternative to using recycled flue gases isto inject steam to control the flame temperature [46]. To furtherincrease the performance of these systems, pressurized systems havebeenproposed for both oxy-coal combustionwith recycledflue gases[47,49e53] and oxy-syngas combustion in combination with solidfuel gasification technology [48]. These approaches are described ingreater detail in the following sub-sections.

1.4.1. Atmospheric oxy-coal combustion systems with flue gasrecycle

The atmospheric oxy-coal combustion system shown in Fig. 4was first introduced as a short-term solution to retrofit existingcoal-fired power plant to include the option of CCS. In most oxy-coal system studies, recycled flue gases at various RR are used tocontrol the flame temperature in the combustor and as a result, theflue gas consists primarily of steamwhich is later removed through

Fig. 4. Atmospheric oxy-coal combustion system with flue gas recycle proposed for carb

condensation, and carbon dioxide which is purified before beingsent for compression and sequestration. The additional equipmentrequired, when compared with air-fired systems, is describedbelow:

1.4.1.1. Air Separation Unit (ASU). When retrofitting existing PCpower plants, the system primarily uses existing equipment withthe exception of an ASU used to produce an oxygen rich stream forcombustion. Currently, the only ASU technology that can meet thevolume and purity demand of a large scale coal-fired utility boiler isbased on cryogenic distillation. Air is compressed, cooled andcleaned prior to being introduced into the distillation column toseparate air into an oxygen-rich stream and a nitrogen-rich stream.Cryogenic air separation is energy intensive, consuming about0.24 kWh/kg O2 with 95% oxygen purity [15,54]. Although theoxygen purity requirement for oxy-coal combustion (85e98%) islower than that needed in the process industry (99.5e99.6%) [55],these cryogenic separation processes can consume more than 15%of the gross power output [15,56e58].

Conventional adsorptionmethods andmembrane separation arealso commercially available, however these methods have not beenapplied to largevolumetric gasproductiondue to thehigher cost andintegration complexity. Other advanced air separation technologiesare still in the inception phase or under development, such as theion-transportmembranes (ITM) andchemical loopingair separation(CLAS). In an ITM, oxygen is separated by exploiting the oxygenpartial pressure gradient across nonporous, mixed-conducting,

on capture in coal power plants, figures are revised based on the work in [20e22].

Page 9: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214164

ceramic membranes at 800e900 �C. If successfully scaled up to CCSapplications such as IGCC, the overall plant efficiency is expected toincrease by1e3percentagepointswith about a 6.5% reductionof theCOE [59]. However, this technologymay not be directly transferableto solid fuel combustion. On the other hand, the CLAS has been latelyproposed for air separation, which relies on circulating metal oxideparticles between an oxidation reactor and a reduction reactor,a chemical principle similar to that used in the CLC [60].

1.4.1.2. Carbon dioxide purification unit (CPU). CPU consists of gascleanupunits to removewater, particulatematter andother pollutantgases from the flue gas before being compressed for sequestration.Because oxy-combustion is compatible with retrofits, selective cata-lytic reduction (SCR), electrostatic precipitator (ESP) and flue gasdesulphurization (FGD) are typically retained as means of NOx,particulate matter and SOx removal from the flue gases. Thesepollutant control devices are also suitable for use in conjunctionwithamine-type absorbents for post-combustion capture plants.

It has been widely accepted that the non-condensable impuri-ties, such as O2, may cause cavitation damage and corrosion in thepipeline during transportation, and this has raised doubts about thesafety of the storage sites. Therefore, after the removal of acid gasessuch as SOx and NOx, non-condensable N2, O2, and Ar should also bepurged using a non-condensable gas purification unit. This unit ismade of multi-stage compression units with inter-stage cooling inorder to separate out the inert gases. Up to the time of this review,there are still no agreed upon standards regarding the requiredpurity of CO2 for storage and sequestration. However, it should benoted that the acceptable degree of purity of the storage-ready CO2results from a trade-off between efficiency losses and operationalcosts during purification and the safety demands of transportationand storage. For a detailed discussion of this topic, the reader isreferred to [20].

1.4.1.3. Flue gas recycle (FGR) system. Recycled flue gas is requiredto replace nitrogen and moderate the combustion temperature.Considering system efficiency and operation practices, flue gasescan be recycled at different locations downstream of the econo-mizer in the form of wet or dry recycles. In the early stages of oxy-coal system studies, the requirement on CO2 purity was not strin-gent and the desulfurization and de-NOx equipment were regardedas unnecessary [61,62]. Therefore, all the flue gas was proposed tobe extracted from a single location downstream of the ESP inwet ordry forms [61]. Later on, Dillon et al. [62] proposed flue gas recy-cling at different locations for the primary (used for transportingcoal) and secondary streams for the sake of energy efficiency: whilethe primary recycle has to be dried and reheated to 250e300 �C totake up moisture from the coal feed, the secondary stream can berecycled at higher temperatures without drying to eliminate ther-modynamic losses caused by cooling and re-heating [62].

Today, with a stricter requirement on CO2 purity for pipelinetransportation and storage, pollution control equipment have beenagain taken into account in the flue gas recycle configurations.Moreover, since SO2 concentration in the flue gas may accumulatedue to flue gas recycle, resulting in 2 or 3 times higher concentra-tion than in conventional air-firing systems, the primary recycle hasto be at least partially desulphurized for medium and high sulfurcoal, to avid corrosion in the coal mill and flue gas pipes.

1.4.2. Pressurized oxy-coal combustion systemsPressurized oxy-fuel combustion systems have been proposed

recently, with the objective of improving the energy efficiency byrecovering the latent heat of steam in the flue gas. The flue gasvolume is reduced under elevated pressure, which results insmaller components and possible reductions in capital cost for the

same power output. Several studies have reported on the technicaland economic feasibility of this process [47,49e53,63,64], allconcluding that the overall process efficiency improves withincreasing operating pressure. This is mainly because latent heatrecovery from the flue gases becomes possible at higher tempera-tures. Other potential advantages of pressurized oxy-fuel systemsare the reduction of the auxiliary power consumption such as therecycle fan work, and the elimination of air ingress into the system.However, there are challenges associated with combustion andheat transfer characteristics at elevated pressures, and hence theburners, steam/gas heat exchangers and condensing heatexchangers must be redesigned [65].

Fig. 5 illustrates two different pressurized oxy-coal combustionsystems proposed in the literature. One of the first designs is theThermoEnergy Integrated Power System(TIPS) proposed and studiedby CANMET [49,66] and Babcock power [53]. This system (Fig. 5a)uses a pressurized combustion unit and heat exchangers, as well asa flue gas condenser (FGC). Downstream of the radiative boiler andconvective heat exchangers, steam in the flue gases is condensed inthe FGC, where most of the latent heat in the flue gas is recovered bythe feedwater in the steam cycle. The rest of the flue gas, which isessentially CO2, is purified and compressed to the sequestrationspecifications. In contrast, in the pressurized system proposed byENEL based on a combustion process patented by ITEA [67e69], andanalyzed by MIT (Fig. 5b) [47,52], the hot flue gases from the pres-surized combustor is quenched to about 800 �C by the recycled coldflue gas, eliminating the need for a radiant heat exchanger and thusincurring a lower capital cost. It should be noted that in these pres-surized oxy-coal systems coal is fed in the form of coal-water slurry(CWS). Since the pressurized system takes advantage of the latentheat recovery from the steam in thefluegas, using a coal-water slurrydoes not significantly decrease the overall energy efficiency.

For the pressurized oxy-fuel power plants with CO2 enrichedflue gas streams, desulphurization and NOx removal solutions havebeen proposed with potentially lower cost and higher energy effi-ciency, using lead chamber chemistry and nitric acid chemistry atelevated pressures. For instance, Air Products [70,71] proposedutilizing two high pressure countercurrent reactive absorptioncolumns (see Fig. 5(b)) while Iloeje et al. [72] combines them intoa single high pressure column to remove SOx as H2SO4 and NOx asHNO3. Both solutions claim to have significantly reduced the cost ofCO2 purification with the latter having an advantage in terms ofreduced power consumption and capital cost.

Although the above-mentioned studies agree on the benefits ofa pressurized system, they differ on the recommended operatingpressure. This difference lies mainly in the design of the flue gascondenser and in disagreement on the amount of recoverablelatent heat at different pressure levels. In the TIPS system studiedby CANMET [66] and Babcock power [53], a temperature differencegreater than 100 �C is maintained between the flue gas and thefeedwater in the flue gas condenser, with the flue gas temperatureat the condenser outlet being 123e262 �C. The high temperaturedifference ensures a sufficiently high heat flux and a relatively smallheat exchanger size. The amount of recoverable latent heat and thenet efficiency increase monotonically with the operating pressure.CANMET suggested that the benefit of increasing the pressurebeyond 80 bar is relatively small, while the Babcock power studyindicated that the benefit of latent heat recovery is mostpronounced between atmospheric pressure and 20.7 bar.

Unlike the TIPS system, the system proposed by ENEL based ona combustion process patented by ITEA [67e69] uses a muchsmaller temperature difference of about 20 �C in the flue gascondenser, with the flue gas temperature at the condenser outletbeing roughly 50 �C. In so doing, most of the latent heat can berecovered at 11 bar [52]. As shown in Fig. 6, the net efficiency for the

Page 10: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Fig. 5. Pressurized oxy-coal combustion systems proposed for carbon capture in coal power plants, figures are revised based on the work in [52,53,66,78]. (a) Schematic of theThermoEnergy integrated power system (TIPS), (b) System proposed by ENEL based on a combustion process patented by ITEA, and analyzed in recent studies by MIT.

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 165

ENEL system reaches a maximum value at a combustor pressure of11 bar, very close to that of the TIPS system at 90 bar. Furtherincreasing the system pressure does not yield additional benefits interms of net efficiency. In order to optimize the equipment size andreduce the capital cost, Gazzino et al. [63] proposed to maintain thesame maximum flue gas velocity in the convective heat exchangerswhile increasing the system pressure. In such a case, for a fixed fluegas mass flow rate, the cross-sectional area of the heat exchanger isreduced, but the length is increased tomatch the total heat transfer,which entails a higher pressure drop. Therefore, increasing theoperating pressure above 11 bar results in a reduction in the net

Fig. 6. Net efficiency of pressurized oxy-coal combustion systems with increasingoperating pressure, simulation results for the TIPS and ENEL system are comparedbased on the work in [52,53]. –B– is for the TIPS system, and —— is for ENEL system.

efficiency. At this point, no studies have been carried out toinvestigate the optimal trade-off between the capital cost (equip-ment size) and the net power efficiency.

Another pressurized oxy-combustion concept that has beensuggested involves the gasification of solid or liquid fuels toproduce gaseous fuel, or syngas. The syngas is combusted in oxygendiluted by recycled CO2 and/or steam, yielding an exhaust gasconsisting mainly of CO2 and/or steam, which is used as a workingfluid in gas turbine. This new configuration and working fluidproperties would require adaption of the current turbomachinerycomponents [73e76]. Anderson et al. [48] applied the oxy-fuelcombustion concept to the burning of syngas, while takingadvantage of a gas turbine (and a combined cycle) to furtherincrease the efficiency, in the so called “Oxy-Fuel Zero EmissionPower Plants” (O-F ZEPP). In their initial design, a GE J79 turbinewas adapted to run as an intermediate pressure turbine connectedto the oxy-syngas combustor exit. The overall efficiency achieved is30e34% based on the lower heating value (LHV) after taking theASU and CO2 compression work into account.

1.4.3. Performance of the oxy-coal combustion systemsAn important question to address at this juncture is the

comparative performance of the atmospheric and pressurized oxy-fuel combustion systems described above. Fig. 6 shows the capitalexpenditure ($/kWe) and efficiency (HHV%) of these systems fornewly-built power plants, compared to the performance of super-critical pulverized coal systems without capture and with post-combustion capture. Data are summarized from independentstudies carried out by NETL [15], MIT [14,47,52], CANMET [66,77],ThermoEnergy [50,78,79], and Kanniche et al. [16]. It is noteworthythat estimates in the open literature vary according to theirassumptions and approximations. For instance, fuel type, size andconfiguration of the power plants, percentage of CO2 captured, andparameters of the steam turbine, etc. Allowing for differences in

Page 11: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Fig. 7. Comparison of plant generating efficiency and capital expenditure[14e16,47,77e79] of CO2 capture technologies. PC: Conventional PC system withoutcapture, post: PC with post capture, A-Oxyf: Atmospheric oxy-coal with flue gasrecycle, P-oxyf: Pressurized oxy-coal with flue gas recycle.

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214166

modeling assumptions, the results from these studies are averagedin Fig. 7, with the minimum and maximum values shown as errorbars; and they should only be compared qualitatively.

System efficiency estimates showed a loss of about 10e15%percentage points when post-combustion capture is added to thebase case PC power plant. On the other hand, the atmosphericoxy-fuel combustion shows an advantage of 1e5 percentagepoints when compared with post-combustion capture; while thepressurized system gains a further 3 percentage points efficiency.The main advantage of pressurized oxy-fuel system is the highersaturation temperature of water at elevated pressures, whichenables more thermal energy recovery and the recuperation oflatent enthalpy, as stated previously. Although the powerconsumption of the ASU is higher in the pressurized combustionsystem, the power savings in the CO2 compression unit and in therecycled flue gas compressor is even higher, culminating ina better overall efficiency [47].

There are significant variations in capital costs estimates inthese studies due to inflation since 2004. The MIT study [14]showed lower cost estimates for the PC without capture, post-combustion, and atmospheric oxy-fuel systems, because it wasbased on the cost of 2000e2004; while Pomalis et al. [77] esti-mated higher costs. The economic studies may only be viewed asrelatively comparable values, but not absolute values before theyare evaluated from commercial scale deployments, especially forthe pressurized oxyfuel systems. It should be noted that these dataare for newly-built power plants, in fact, the capital cost estimatesfor retrofitted atmospheric oxy-coal power plants are significantlylower than any of these options as discussed in Section 1.1.

Table 3Comparison of selected physical properties of CO2 and N2 at 1 atm and 1000 K. Data are

Physical process Physical property Un

Thermodynamic Density kgSpecific heat capacity kJVolumetric heat capacity kJGas-water interfacial tensiona N

Momentum transfer Kinematic viscosity mHeat transfer Thermal conductivity W

Thermal diffusivity mAbsorptivity/emissivity

Mass transfer Mass diffusivityb m

a Water gas interfacial tensions are evaluated for CO2/water and air/water interfaces ab Mass diffusivity refers to the binary diffusion of O2 in CO2 and nitrogen.

2. Thermodynamics, transport and chemistry in oxy-coalcombustion

This section focuses on the thermodynamics, transport andchemistry processes that take place during oxy-fuel combustion ofcoal. Experimental and numerical studies of single coal particles ora group of coal particles are reviewed with emphasis on the impactof a CO2-rich environment. We start with a discussion on the heattransfer characteristics in oxy-fuel combustion and how to matchthe heat flux with that in the conventional air-fuel combustion,followed by a description of the stages and processes that a coalparticle undergoes as it is entrained into the furnace: trans-portation, heating and moisture evaporation, devolatilization,ignition and burning of the volatile matter, flame propagation, charcombustion, and pollutant formation. The physical properties of N2and CO2 relevant to oxy-fuel combustion are summarized inTable 3. The effects of these properties will be discussed in greaterdetail in each of the following sub-sections.

2.1. Heat transfer

Extensive studies have been conducted on retrofitting existingcoal-fired power plants for oxy-fuel combustion. Most studiesagree that the first priority in the retrofit effort is to maintain thesame heat transfer characteristics in the furnace as in air-firedcombustion. Matching the heat transfer in the furnace can lead tosimilar combustion stability, carbon burnout, and slagging andfouling tendencies [33]. The following subsections focus on thefundamentals of radiative and convective heat transfer in oxy-fuelcombustion, and discuss methods to maintain similar heat trans-fer characteristics when retrofitting a PC boiler to an oxy-firedmode, based on a review of the experimental and simulationstudies on temperature distribution and heat transfer in furnacesand boilers.

2.1.1. Radiative heat transferRadiative heat transfer plays an important role in boilers and

furnaces. The characteristics of radiative heat transfer in oxy-combustion are distinct from conventional air-coal combustionbecause the flue gas compositions are different. A simplified scalinganalysis can be used to estimate the radiative heat transfer ratefrom an oxy-fuel combustion flame. The entire flame is consideredto be a uniform source of radiation, and its total radiative energyrelease rate can be approximated as:

_qradfεAf sT4f (2)

where ε is the average emissivity for the flame, and Af and Tf arethe flame surface and temperature, respectively. The flametemperature is the overwhelming factor because of its 4th orderdependence [80]. Therefore, to match the radiative heat flux

cited from [99,164,245,246].

it CO2 N2 CO2/N2

=m3 0.5362 0.3413 1.57=kgK 1.2343 1.1674 1.06=m3K 0.662 0.398 1.66=m 71.03 71.98 0.9872=s 7.69e-5 1.2e-4 0.631=mK 7.057e-2 6.599e-2 1.072=s 1.1e-4 1.7e-4 0.644

>0 w0 e2=s 9.8e-5 1.3e-4 0.778

t 298.15 K.

Page 12: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 167

between air and oxy combustion, one must achieve gas tempera-tures in the latter that are close to those found in the former. On theother hand, for the same flame temperature, radiation heat transferis enhanced when the emissivity is higher. Unlike symmetricdiatomic gases such as N2, triatomic gases such as CO2 and H2O arenot transparent to radiation. Their partial pressures are signifi-cantly higher in oxy-fuel combustion flue gas than those in air-fuelcombustion, and correspondingly, the absorptivity and emissivityof the flue gas substantially increases.

When a radiation beam of specific wavelength l travels througha gas, a portion of the radiation intensity, al, is absorbed by the gas.The change in the spectral radiation intensity, dIl within an infin-itesimal beam length ds is:

dIl ¼ �klIlds (3)

where kl is the spectral absorption coefficient, which is a func-tion of the wavelength, gas composition, gas temperature andpressure [81]. The spectral absorptivity and emissivity can bederived by integrating the above Eq. (3) over the beam length, s:

al ¼ εl ¼ 1� exp½�kls� (4)

Fig. 8 shows the spectral absorptivity, al, of the different gascomponents produced in a combustion process [82]. Sincetriatomic gases (CO2 and H2O) have much higher partial pressureswhen flue gases are used as the diluent instead of nitrogen, theabsorption and radiation in oxy-combustion are stronger than inconventional air-fuel combustion with identical gas temperatures.

A close look at Fig. 8 shows that absorption is not continuousover the entire spectrum; instead it is concentrated in a number ofmoderately wide spectral bands. The effective absorptivity andemissivity of a gas mixture can be obtained by integrating theabove equation over the full spectrum. However, direct computa-tion of the gas absorptivity is difficult, and gray gas or band modelsare widely used in the CFD modeling, which will be discussed inSection 4.3.

Soot and particle (coal, char and fly ash) radiative emission andabsorption play as important a role as the triatomic gases in coalcombustion [83,84]. Experimental studies in oxyfuel combustionhave confirmed this trend. Andersson and coworkers [85e88]investigated the radiation intensity of propane-fired and lignitecoal-fired oxy-fuel flames in the Chalmers’100 kW test facility. Thetotal radiation intensity was measured using a narrow angle radi-ometer, and the gas radiation was estimated using MalkmusStatistical Narrow BandModel (SNBM). Results of the calculated gas

Fig. 8. Spectral absorptivity as a function of wavenumber for water vapor, carbondioxide and methane, reproduced from reference [82].

and the measured total radiation intensities at the flame zone areshown in Fig. 9. It can be seen that the portion of the gas radiationin the total radiation intensities under OF25 condition (OF25denotes oxy-fuel combustion with O2 concentration of 25% in theburner gas) is significantly higher than in air-combustion.Measurements shows that the temperature distribution in theflame zone under OF25 condition is very close to that under the air-firing condition (see Section 2.1.3 for greater detail), and that theH2O partial pressure increases only slightly from air to OF25condition, therefore, the higher gas radiation in oxy-fuel combus-tion is attributed to its 5 times higher partial pressures of CO2.However, the effect of the enhanced gas radiation during oxy-fuelcombustion becomes less important when comparing to theparticle radiation; the particle radiation contributes about 60e70%of the total radiation in both air-fired and oxy-fired cases. Thisstudy suggests that in lignite oxy-fuel combustion, if the gastemperature is maintained the same as in air combustion byincreasing the O2 concentration to 25% (OF25), the measured totalradiation intensities are similar (OF25-total and Air-total) althoughgas radiation is enhanced. This is because a large fraction of radi-ation is emitted by particles, which have similar contribution inboth combustion environments.

2.1.2. Convective heat transferThe convective heat flux of the flue gas, q00conv, can be approxi-

mated as follows:

q00conv ¼ hDT (5)

where h is the convective heat transfer coefficient, and DT is thetemperature difference between the bulk gas and the heated object.The convective heat transfer coefficient is influenced by the flowvelocity and gas properties such as viscosity, thermal conductivity,heat capacity and density, which are also functions of temperature.The ratioof convectiveheat transfer coefficients of thefluegas inoxy-fuel combustion (hoxy) to that in air-fuel combustion (hair) can beexpressed in termsof dimensionless numbers - theReynoldsnumberand the Prandtl number - and the fluid thermal conductivity [33]:

hoxyhair

¼�ReoxyReair

�m�ProxyPrair

�n�koxykair

�(6)

where m and n are empirical factors that vary for differentgeometries. The slightly higher thermal conductivity of CO2 does

Fig. 9. Comparison of total intensity measurements (symbols) and gas radiationmodeling (lines) at 384 mm away from the burner inlet. Data cited from Anderssonet al. [86].

Page 13: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Fig. 11. The O2 partial pressure (fraction) required at burner inlet (to achieve similaradiabatic flame temperature as the air-fired case) for wet and dry flue gas recycle(residual O2 mole fraction in the flue gas fixed at 3.3%) [22]. The symbol - indicatesthe AFT of air-coal combustion, the red solid line and blue dash line indicate theAFT of oxy-coal combustion with dry and wet flue gas recycle, respectively. (Forinterpretation of the references to colour in this figure legend, the reader is referred tothe web version of this article.)

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214168

not significantly change the conductive heat transfer whencompared with air-combustion. However, for convective heattransfer at identical velocity, its lower kinematic viscosity results ina larger Reynolds number and thus a higher convective heattransfer coefficient.

Woycenko et al. [89,90] studied convective heat transfer coef-ficients in oxy-coal and air-coal combustion. The ratio hoxy=hair isshown in Fig. 10. This ratio increases with increasing recycle ratiobecause of the changes in the thermal properties of the flue gas andthe increase in the gas velocity. This figure also shows that, for therecycle ratios reviewed in this paper, the convective heat transfercoefficient ratio is slightly above unity in most of the cases. Forinstance, this value is w1.15 when the wet recycle ratio is 70%, atwhich the adiabatic flame temperatures are approximately thesame.

Recall that in Eq. (5), the convective heat flux is a function of thetemperature difference (DT), which on average decreases withincreasing recycle ratio. Therefore, an acceptable operational rangeof the flue gas recycle ratio exists in which one can approximatelymatch the convective heat transfer in oxy-coal combustion to thatof air combustion [20,91]. Thus the existing coal-fired power plantcan be retrofitted to function as an oxy-coal plant without replacingthe boiler or heat exchangers.

2.1.3. Matching of combustion temperatureThere are several constraints that need to be considered when

retrofitting a conventional air-fired power plant to oxyfuelcombustion, including the combustion temperature, the furnaceoutlet temperature, and the heat transfer in radiative and convec-tive heat exchangers. Based on the above analyses, it appears thatmatching the gas temperature profiles in the boiler is a startingpoint in the effort to maintain similar heat transfer characteristicsin oxy-coal combustion. A good indication that the temperatureprofiles are matched is when a similar adiabatic flame temperature(AFT) is achieved. Wall et al. [22] calculated the theoretical O2 molefraction (or flue gas recycle ratio) necessary to maintain the sameAFT under air-coal and oxy-coal combustion conditions using theequilibrium approximation. Approximately 20% excess air in air-coal combustion and 3e5% excess O2 in oxy-coal combustion areused to ensure that all the calculations are done on the same basisof about 3.3% (v/v) residual O2 in the flue gas. Fig. 11 shows thecomputational results of the AFT in air-fired and oxy-fired condi-tions with increasing oxygen fractions for both wet and dry flue gas

Fig. 10. Effect of recycle ratio on convective heat transfer coefficient [89]. Primary O2 isthe oxygen mole fraction in the primary stream.

recycle. The results suggest a temperature drop of about 400 K and800 K, respectively, in wet and dry recycling oxy-fuel combustionwith oxygen concentration of 21% by volume when compared withthat of the air-fired condition. The authors did not indicate the coaltype in their calculation; it should be noted that the results aredepend on the approximate and ultimate analysis of coal.

Table 3 summarizes selected physical properties of CO2 and N2and their ratios at 1000 K and 1 atm in four categories: thermo-dynamic properties, momentum transfer properties, heat transferproperties and mass transfer properties. From the data shown inTable 3, although the specific heats of CO2 and N2 are similar onmass basis, the CO2 heat capacity on mole basis is 1.66 times higherthan that of N2 because of its larger molecular weight. Since the N2is replaced with CO2 and H2O under wet recycle oxy-fired condi-tions, and with CO2 in dry recycle oxy-fired conditions, there isa significant reduction in the combustion temperature. In order tomatch the AFT with that of air-fired combustion, the overall oxygenmole fraction in the burner inlet should be increased to about 28%or 35% for wet and dry recycle, respectively, to compensate for theeffects of the higher molar heat capacities of H2O and CO2. It is alsonoteworthy that in practice, the combustion temperature cannot beas high as predicted by the equilibrium calculations carried out byWall et al. [22] because of the heat absorbed by the radiative heatexchangers.

The gas phase temperature profiles in the furnace have alsobeen measured experimentally to investigate the influence of theflue gas recycle ratios on the gas temperature and heat transfer inthe furnace under oxy-coal combustion conditions, with theobjective of maintaining a heat flux similar to that observed in aircombustion. Selected researches are listed in Table 4. Most of thesepilot scale facilities have refractory walls lined with water jackets tosimulate the combustion conditions in utility boilers. However, theheat transfer characteristics of these test facilities may vary due tofacility setup and operating conditions. The measured temperatureprofiles are given at limited locations (usually in the centerlines ofthe furnaces), and are subject to measurement errors from radia-tion effects, and finite response times. Therefore, they can only beused as general estimates of the gas temperature distribution in thecombustion zones.

Page 14: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Table 4Bench and pilot scale experimental studies on gas temperature and heat transfer in atmospheric oxy-coal combustion.

Organization Fuel Facility Recycle mode Experiment findings

Argonne NationalLab/EERCa [31,33]

Black ThunderSub-bituminousCoal from Wyoming

3 MWth, vertical tower furnace Wet/Dry FGR Flue gas recycle ratios ((CO2þH2O)/O2

in the oxidant) of 3.25 and 2.6, in wetand dry recycle respectively, matchesthe overall heat transfer of the air-firedfurnace

Argonne NationalLab/BCL [32,33,250]

Wage coal from Colorado 117 kWth, bench scalefurnace, i.d. 0.6 m

CO2/O2 mixture Dry recycling with a CO2/O2 molar ratio of 2.23yields a furnace temperature distributionsimilar to that of air combustion.

IFRF [89,90] Coal 2.5 MW, 2*2 m internalsquare cross-section

Wet FGR 58e61% of the flue gas is recycled for stablecombustion and matched convective heattransfer coefficient and radiative heat flux

University ofLeeds [182,183]

Bituminous coals fromUK and internationalsources

Down-fired 20 kWthcombustor

CO2/O2 mixture When air is used as the primary stream,an O2 concentration of 30% in the secondarystream produces matching gas temperatureprofiles. When an oxygen concentration of30% is used in both primary and secondarystream, the resulting gas temperature is higherthan air-coal combustion.

Chalmers Universityof Technology [85e88]

Propane and lignite coal 100 kWth Dry FGR Oxy-Propane combustion: With an O2 molefraction of 25%, the temperature profiles inthe furnace can be matched with that ofair-propane combustion. Soot formation alsoplays an important role in the radiationintensity in air and oxy-combustion ofpropane. Calculated gas radiation is higherin oxy-combustion than in air combustion,but contributes only 30%e50% of the totalradiation intensity.Oxy-Lignite coal combustion: With an O2

mole fraction of 25%, the temperaturedistribution can be matched with that ofair-fuel combustion; the flue gas temperaturesat the furnace outlet are lower in all theoxy-fuel cases than in air-fuel case.

IHI [95] Low- and medium-volatilebituminous Coal

1.2 MWth horizontalcombustion test facility

Wet/Dry FGR When using a secondary gas stream withan oxygen mole fraction of 30%, and dryrecycled flue gas as the primary gas, the gastemperature near the burner is lower inoxy-fuel combustion; the gas temperatureis greatly increased with direct oxygeninjection; drying the primary gas increasesthe gas temperature by 150 �C near the burner.

RWEn power [91] Russian and SouthAfrican coals

0.5 MWth, RWEncombustion test facility

Simulated Dry FGR Peak radiative heat flux values are inverselyrelated to the recycle ratio. Conversely,the convective heat transfer values increasewith increasing recycle ratio. A matchedradiative heat transfer is established atrecycle ratio between 68% and 72% forRussian coal, and between 72% and 75% forSouth African coal, respectively.The experimental data show that matchedheat transfer characteristics can be establishedwith about 74% dry recycle ratio for Russian coal.

CCSD/IHI [22] Three Australian coals 1.2 MWth pilot scalevertical test facility

Wet FGR When using an average oxygen mole fractionof 27% in the burner streams, the maximumflame temperature in oxy-fuel combustion is100e150 �C lower than the air-fuel flametemperature. The gas temperatures downstreamof the burner are almost matched with that ofair-fuel combustion.

E.ON UK/Hitachi [189] Tselentis coal 1 MWth, horizontal firing Dry FGR Gas temperatures of the oxy-fuel flame witha dry recycle ratio of 76% (21.3% O2 in theburner gas) are lower than the temperaturesin air-fired flame.

IVD/Hitachi [189] Lausitz lignite coalfrom Germany

500 kWth, down-shot firing Wet FGR Gas temperatures of the oxy-fuel flame with awet recycle ratio of 53% (39.3% O2 in theburner gas) are lower before the level 3measurement port, and higher at allfollowing levels.

CANMET [92,94] Bituminous,sub-bituminous,and lignite coals

0.3 MWth, vertical combustorresearch facility (VCRF), cylinderdown-flow, 0.6 m inner diameter

CO2/O2 mixture,or Partial FGR withmake-up CO2

With an O2 mole fraction of 28%e35%,the gas temperatures and heat fluxescan be matched with those of the air-coalcombustion.

a Energy and Environment Research Corporation in Irvine, CA.

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 169

Page 15: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Fig. 13. Comparison between wet and dry recycles of primary gas [95].

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214170

Awide range of oxygenmole fractions (about 23.5e35%) neededto achieve gas temperatures and heat transfer performances similarto those in air-coal combustion have been reported. For instance,CANMET [92e94] conducted a series of experiments in a 0.3 MWvertical combustor research facility (VCRF) to investigate theinfluence of oxygen mole fraction on the combustion temperatureand heat transfer in the furnace. A broad range of oxygen concen-trations, from 21% to 42%, was tested in O2/CO2 modes with bottleCO2 as well as in air-fired conditions [92]. The gas temperatures andheat fluxes were found to match air-combustion with O2 molefraction between 28% and 35% in oxy-fired conditions for bitumi-nous and sub-bituminous coals. Similar results were obtainedwhen using partial flue gas recycle [94]. Andersson et al. [86] andHjartstam et al. [88] needed a relatively lower oxygen mole fraction(about 25% in the feed gas) to achieve a gas temperature equivalentto that of air-lignite coal combustion (see Fig. 12). When operatingwith higher oxygen mole fractions of 27% and 29%, they found thatcombustion temperatures are 50 �C and 100 �C higher than that ofair-coal combustion, respectively.

One of the most important factors for determining the oxygenlevel needed to match the temperature field is whether the recy-cled flue gas is wet or dry. This is because of the higher volumetricheat capacity of CO2 as compared with H2O. Compared with Wall’scalculation [22] in which O2 mole fractions of 28% and 35% arerequired, respectively, to maintain the theoretical AFT in wet anddry recycle conditions, experimental results in the literature, listedin Table 4, indicate a slightly lower mole fraction: 23.5%e27% forwet recycle, and 25%e35% for dry recycle. This may be because ofthe lower combustion temperatures encountered in the practicalfurnaces. Moreover, no explicit relation between the oxygen molefraction and the scale of the tested furnace thermal capacity hasbeen found in the experimental studies shown in Table 4.

An exception to the conclusions stated above is reported in theexperimental findings of Nozaki et al. [95], who performed oxy-coalcombustion experiments on the IHI 1.2 MW horizontal combustiontest facility using a low- and a medium-volatile bituminous coal.When using wet flue gas in the primary stream and an oxygenmolefraction of 30% in the secondary stream, the combustion gastemperature near the burner was significantly lower than that ofair-coal combustion. Drying the flue gas in the primary streamwasfound to be beneficial to the combustion temperature. Fig. 13 showsthat drying the primary gas helps to increase the combustiontemperature by about 150 �C. However, the paper did not explicitlystate whether the recycled flue gas volumetric flow rate was fixed

a b

Fig. 12. Measured radial gas temperature profiles for air-lignite and oxy-lignitecombustion (a) 384 mm and (b) 800 mm from the burner inlet [86].

in the case with dry flue gas. If the recycled flue gas volumetric ratewas fixed, the recycled stream flow rate to the burner would havebecome lower because water vapor was removed, and the flametemperature could have been higher because of the lower totalrecycled gas volume.

The optimal oxygen fraction for a matched temperature is alsodependent on the fuel, and this dependence needs further inves-tigations. Fig. 14 shows the measured temperature profiles in theCANMET bench scale vertical combustor research facility [92]. Itcan be seen that the gas temperature of the Canadian western sub-bituminous coal (Highvale) oxy-combustion with an oxygen molefraction of 35% is 80 �C higher than in air-coal combustion, whereasthe temperature difference between the two combustion envi-ronments with the same oxygen mole fraction is much smaller forthe higher rank US Eastern bituminous coal. In other words, therequired oxygen mole fraction for the higher rank Eastern bitumi-nous coal was higher. To the contrary, Smart et al. [91] conductedradiative flux measurements in the RWEn 0.5 MWth combustiontest facility firing two different coals under oxy-fuel conditions andobserved the opposite trend. Their results, shown in Fig. 15,compare the radiative flux in the furnace when firing Russian andSouth African coals. It can be seen that a recycle ratio between 68%and 72% produces a heat flux profile similar to air firing for the

a b

Fig. 14. Temperature profiles along the centerline in oxy-coal combustion: (a) Highvalesub-bituminous coal and (b) Eastern bituminous coal [92].

Page 16: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

a b

Fig. 15. Axial radiative heat fluxes in air-coal and oxy-coal combustion with increasingrecycle ratios. (a) Russian coal and (b) South African Coal. All the tests are operatedwith dry recycle and 3% residual O2 [91].

a b

Fig. 16. Measured heat extraction distribution in a 3 MWth pilot scale boiler in (a) Dryrecycle and (b) wet recycle [33]. Plots show the heat extraction at different heatexchanger sections. Solid symbols in the vertical axis indicate those under air-coalcombustion condition.

Fig. 17. Optimum recycle ratios for a matched heat transfer in the furnace. Comparisonof predictions (denoted by lines) for bench-scale and pilot-scale furnaces (measureddata denoted by B, ,) and utility boiler (only show predictions) [33].

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 171

Russian coal, while the ratio is between 72% and 75% for the higherrank South African coal. This difference in recycle ratios indicatesthat the required oxygen mole fraction for the higher rank SouthAfrican coal is lower. Based on these limited findings, no conclu-sions can be drawn concerning the dependence of oxygen molefraction, or recycle ratio, on the coal type (such as its rank, moisturecontent, etc.), and further investigations are required to addressthis topic.

2.1.4. Matching of radiative and convective heat transferdistribution

Considering that the radiative and convective heat transfercharacteristics in oxy-fuel combustion are rather different fromthose in air-fuel combustion, a matched gas temperature might beclose to, but not precisely, the optimum operating condition inorder to obtain a matched distribution of heat fluxes in differentheat exchanger components. Hence, studies have also been con-ducted in order to find the optimum recycle ratios for a matchedheat transfer distribution in the furnace (radiative section) and inthe convective section. Based on the discussion in Section 2.1.1,radiation dominates the heat transfer process in the furnace.Radiation intensifies with increasing flame temperature anddecreasing flue gas recycle ratio. Therefore, to match the heattransfer distribution between radiative heat exchangers for phasechange and convective heat exchangers for superheat, there is anoptimal flue gas recycle ratio. Lower recycle ratios result in highercombustion temperatures and excess heat extraction in the radia-tive section, especially in the water-cooling membrane section. Inthis case, the required superheat and reheat temperatures of thesteam cannot be attained. Conversely, higher recycle ratios result ina lower combustion temperature and a reduced steam mass flowrate that is overheated. In this case, considerable attemperation (ameasure controlling the overheated steam temperature) withfeedwater injection is required to reduce the steam superheattemperature, and this would reduce the boiler efficiency [33].

In experiments conducted by Payne et al. [33] on a 3MWth pilotscale test furnace, heat extraction in various boiler sections wasmeasured under oxy-coal combustion for a range of recycle ratios,in both wet and dry flue gas recycle modes. As shown in Fig. 16, theoptimal oxygen mole fractions at which the radiative heatexchangers reach heat extractions similar to those of air-firedcondition are 23.5% and 27.8% (corresponding to (CO2 þ H2O)/O2mole ratios of approximately 3.25 and 2.6) for wet and dry recycle,respectively). Under such conditions, the furnace temperatures for

both the wet and dry recycle processes were slightly lower than thecase of air-coal combustion and this is compensated by theenhanced radiation and convective heat transfer in oxy-combustion.

Moreover, the same study [33] showed that the recycle ratio andoxygen mole fraction needed to match the heat transfer distribu-tion of the corresponding air-fired system are also dependent onthe scale of the combustion facility. Fig. 17 shows that the optimalCO2/O2 ratios in dry recycle needed to match the heat transferdistribution are 2.45, 2.6 and 2.7 for 117 kWth bench scale furnace,3 MWth pilot scale furnace and 44 MWe utility boiler, respectively.This scale-dependence is closely tied to the radiative heat transfercharacteristics: the gas emissivity increases with increasing beamlength in larger scale boilers, hence lower flame temperatures arerequired to maintain a similar radiative heat flux, which areattained by a higher recycle ratio.

Similarly, a recent experimental study on a 0.5 MWth RWEncombustion test facilityconductedbySmart et al. [91,96] showed theeffects of varying the recycle ratio on the radiative and convectiveheat flux under simulated dry recycle conditions. Fig. 18 shows

Page 17: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Fig. 18. Measured peak radiative heat fluxes, convective heat transfer coefficients andcalculated adiabatic flame temperature in oxy-fired conditions normalized to those ofair-fired conditions. Operating conditions: Russian coal, simulated dry recycle [91].

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214172

a similar trend as described above, that the radiative heat fluxdecreases with increasing recycle ratio, and the matching radiativeheat flux is attained at a recycle ratio of about 74% with an adiabaticflame temperature that is lower than that in the air-fired condition.The authors suggested that the higher flame emissivity, due to sootand possibly a higher concentration of CO2, compensates for thelower combustion temperature in oxy-fuel combustion. On theother hand, as shown in Fig. 18, the convective heat transfer coeffi-cient (HTC) rises with increasing recycle ratio within the range ofoperating conditions analyzed in the study, and reaches a value thatmatches the HTC under air-fired condition at a recycle ratio of 74%.This trend, caused by the increased mass flow over the convectivetube at higher recycle ratios, agrees with the results of the experi-ments conducted by Woycenko et al. [89] discussed in Fig. 10.

It should be noted that although the HTC matches that observedunder air-fired condition at the same recycle ratio in which theradiative heat flux is also matched (74%), the correspondingconvective heat flux may not match because of the lower flue gastemperature (also see Eq. (5)). Therefore, measures for modifyingthe heat exchangers, such as the refractory lining and/or convectiveheat exchanger cluster adjustment, might be necessary when ret-rofitting existing PC boiler to operate under oxy-combustionconditions.

2.2. Fuel delivery and injection

2.2.1. Pulverized coal transportPulverized coal is delivered to coal boilers using a carrier fluid. In

conventional atmospheric pressure PC boilers, pulverized coal istransported by preheated primary air [97]. In an air-like oxy-coalsystem, recycled flue gases (primarily CO2) are used to transportpulverized coal into the boiler. The lower oxygen concentration inthe carrier gas reduces the risk of coal-dust explosions. While thepneumatic transport of pulverized coal by air was extensivelyinvestigated in the past, similar studies using CO2 as the carrierfluid have yet to be performed. This is worth mentioning becausethe physical properties, for instance the density and the viscosity, ofair and CO2 are different.

To maintain a similar transport performance, the solid-gas massfraction, the carrier gas velocity, and the pressure drop will have tochange. The pick-up velocity, vpu, is defined as the mean velocity ofa gas stream required to transport solid particles in suspension

through a horizontal pipeline without allowing them to settle onthe bottom of the pipe, and it is the minimum conveying velocity inthe design of pneumatic transport systems. Cabrejos and Klinzing[98] carried out experiments to determine the pick-up velocity andfound that the pick-up velocity is roughly inversely proportional tosquare root of the gas density:

vpufr�1=2g (7)

Based on dimensional analysis and experimental data, theyfurther suggested the following correlation for the pick-up velocity,vpu, to convey particles that are larger than 100 mm in a horizontalduct:

vpuffiffiffiffiffiffiffiffigdp

p ¼�0:0428Re0:175p

��Ddp

�0:25 rprg

!0:75

(8)

where g is gravity, dp is the particle diameter, D is the pipediameter, and r, rp are the density of the gas and particle.

Rep ¼ rgvgdpmg

is the particle Reynolds number using mean gas

velocity. This correlation is valid for 25 < Rep < 5000;

8 <Ddp

< 1340; and 700 <rprg

< 4240. Assuming the primary

stream temperature is 90 �C and using the physical propertiesreferred from the NIST database [99], the estimated pick-upvelocities for 1000 mm coal particle in a 0.2 m diameter hori-zontal duct are 9.6 m/s and 7.4 m/s in air and CO2, respectively.

Therefore, when a dry recycle flue gas composed mainly of CO2is used for the pneumatic transport of pulverized coal instead of air,the pick-up velocity is smaller. The fact that the flow velocityrequired to entrain the coal powder is reduced has been observedexperimentally by Cabrejos and Klinzing [98], and Villareal andKlinzing [100]. In a conventional air-fired system, it is typical tochoose an air velocity higher than 20 m/s to avoid the saltation andsettling of solids in horizontal pipelines. Hence, from the aboveargument in Eq. (7), roughly a CO2 velocity of at least 16.3 m/sshould be chosen. However, pneumatic conveying technology isheavily based on engineering experience and experimental studies[101], and more work is needed to determine the guidelines, suchas the minimum carrier gas velocity and pressure drop, when usingrecycled flue gas as the transport fluid.

2.2.2. Coal-water slurryAt higher operating pressures, pulverized coal is typically

transported to the burner/combustor in the form of dry coal by lockhoppers, or in the form of slurry (coal-water slurry (CWS) or coal-liquid CO2 slurry). The dry feed lock hopper system discharges coalparticles into the combustor via staged opening and closing ofvalves on the top and bottom of the pressure vessel. The system canwork under pressures of up to about 40e50 bar [102]. CWS isprepared bywet grinding the coal in a rodmill, which is then storedand pumped into the combustor [103]. The CWS system can beoperated under pressures of up to 70 bar, thus it provides a broaderoperating range for pressurized oxy-coal system. Later studies haveshown that coal-liquid CO2 slurries have advantages over coal-water slurries, leading to higher cold gas efficiencies in gasifica-tion applications [104,105]. Since both the TIPS and ENEL pressur-ized oxy-coal systems use the CWS as the fuel delivery method, weconfine our discussion to this form of coal delivery and review itsatomization characteristics.

The coal-water slurry is atomized into small droplets duringinjection to maximize the reaction surface of the fuel and to facil-itate its mixing with the oxidizer streams before combustion.Several types of atomizers have been developed for this purpose,

Page 18: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 173

such as pressure nozzles, rotary atomizers, twin-fluid atomizers,air-assist atomizers and effervescent atomizers. A review of theoperating principles of these atomizers and their application incombustion systems was given by Beer and Chigier [106].

Sojka and Lefebvre [107] conducted experiments on a CWSeffervescent atomizer, and developed an empirical correlation toestimate the Sauter mean diameter (SMD) as function of the coalparticle size, the atomization pressure, and the GLR. For theatomizer geometry and coal particle size in their study, they foundthe following correlation:

SMD ¼ 12s

rl�vl � vg

�2�1þ 1

εGLR

�(9)

where vl and vg are the estimated velocity of liquid and gasdownstream of the injector before atomization, GLR is the gas toliquid mass flow rate ratio, and ε is the fraction of gas that directlyparticipates in the atomization process, which can be obtainedfrom correlations from experiments. The Weber number is widelyused in atomization and it is defined as:

We ¼ rdv2

s(10)

Eq. (9) and Eq. (10) indicate that the SMD is a function of GLR, εandWeber number.We is the ratio of the kinetic energy of a dropletissuing from the jet relative to its surface energy. A large We indi-cates that the injected slurry liquid sheet has a relatively higherkinetic energy, which breaks up the liquid sheet to form manysmaller droplets. Since the coal-water droplets contain 35e40%water by weight, the gas-water interfacial surface tension is used tocalculate the We of the droplets.

It is interesting to compare the atomization phenomena of coal-water slurry issuing into a gaseous environment dominated bynitrogen, as in air-fuel combustion, with one dominated by CO2, asin oxy-fuel combustion. Fig. 19 shows that the We in a high CO2

concentration stream does not differ much from the We of thedroplets in pure N2 streams at atmospheric pressure. However, thedifference is markedly obvious at elevated pressures (>5 MPa)where the We in the high CO2 concentration stream is significantlylarger. This is mainly because of the smaller interfacial tension ofCO2-water that is observed in experiments [108]. Higher We

Fig. 19. CalculatedWeber number of water droplets (characteristic diameter is 100 mm,relative velocity is 10 m/s) in high CO2 concentration gas and pure N2, at elevatedpressures. Interfacial tension values used are based on [108], the data for 0.1 MPa caseare taken from [245,246]. , indicates the We number in pure N2, and B indicates theWe number in CO2-rich atmosphere.

corresponds to smaller atomization droplet size, indicating thatCWS atomization can be better in pressurized oxy-coal combustion.

2.3. Heating and moisture evaporation of coal particles

Grinded pulverized coal particles are heated up and partiallydried by the primary air or recycled flue gas in the coal mill in orderto avoid blockage of the primary pipe. To ensure drying capacity,the primary air or recycled flue gas should be reheated to above250 �C, and maintained around 60e90 �C at the mill exit [62]. Theremaining moisture content of the coal particle, or the watercontent in the case of CWS fuel, is evaporated in the furnace beforeheating, devolatilization and burning take place. Considering thedifferent heat transfer characteristics in oxy-fuel combustion,heating andmoisture evaporation processes of a single coal particlemight be different, which in turn may influence the heating rate,devolatilization rate and yield, as well as the standoff distance ofthe flame (also denoted as lift-off distance, which is defined as thedistance between the burner tip and the visible ignition of theflame).

In this work, the impact of the gas composition on single coalparticle drying and heating processes is analyzed using a simplifiedheat transfer model, and the evaporation time and heating rate inair-fuel and oxy-fuel conditions are estimated. Fig. 20 shows anillustration of a quasi-steady model that considers heat transferbetween a single coal particle, the surrounding gases and theluminous flame. The energy conservation of a coal particle can beexpressed as the balance of the particle internal energy andconvective/radiative heat transfer:

43pr3prpcp

dTpdt

¼ 4pr2ph�TN � Tp

�þ 4pr2ps�εf T

4f � εpT4p

�(11)

where TN is the temperature of the surrounding hot gas, and Tfis the flame temperature. The heating rate is strongly determinedby the temperatures of the gas and the flame. If the oxygenconcentration is kept the same, the flame temperature for oxy-combustion is comparatively lower than that of the air-fuelcombustion (also see 2.1.3), resulting in a lower heating rate. Thiseffect is considered in the estimation by assigning differentsurrounding gas temperatures. On the other hand, the convectiveheat transfer coefficient of the flue gas, h, may be different, whichalso may result in a different heating rate and evaporation time inoxy-fuel combustion.

Water evaporation from the coal particle is a process thatcouples heat and mass transfer. Using the quasi-steady model of

Fig. 20. Schematic diagram of heat transfer to a single particle in coal combustion.

Page 19: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Fig. 21. Predicted coal particle drying time as a function of particle size in air and CO2

gas atmospheres. Drying processes in the primary duct during fuel transportation andin the furnace before combustion are estimated. The primary gas stream is set at105 �C, and the gas and flame temperatures in the furnace are set to 1000 �C and1800 �C, respectively.

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214174

droplet evaporation [109] and the model of coal-water slurryevaporation proposed by Cen et al. [110], evaporation times fordifferent particle sizes under both air and CO2 conditions arecalculated, with the following assumptions:

1. Coal particles have a spherical shape with a fixed radius duringvaporization; only the density decreases until it reaches the drycoal particle density.

2. The relative velocity between the coal particle and thesurrounding gas is the terminal velocity, vt , given by

vt ¼ d2pðrp � rgÞg18mg

when Rep < 2 [111].

3. The evaporation rate is controlled by heat transfer rather thanmass transfer in the mill and in combustion [109].

4. The contribution of convective and radiative heat transfer isconsidered [110,112]. Research has found that for small parti-cles in PC combustion, convective heat transfer is more domi-nant over radiative heat transfer during heating andevaporation processes [110]. This is because of the small surfacearea of the coal particle which limits the contribution of radi-ative heat transfer. Therefore, in the following analysis, thesame flame emissivity is used to estimate radiative heattransfer, neglecting the difference between flame emissivity inair-fired and oxy-fired conditions.

Using the above assumptions, the evaporation rate ( _m; kg=s)can be derived as:

_m4pr2p

¼ kgNu2cpgrp

ln�1þ Bq

�(12)

where

Bq ¼ cpgðTN � TboilÞ

hfg �4pr2ps

�εf T4f � εpT4boil

�_m

(13)

and the Nusselt number is given by Ranz and Marshall [113] as:

Nu ¼ 2þ 0:6Re1=2p Pr1=3 (14)

The evaporation time is obtained by integrating Eq. (12):

sboil ¼2ðr0 � rcoalÞCpgr2p3kgNuln

�1þ Bq

� (15)

The drying process in the mill and the furnace, and the heatingprocess after evaporation under both air- and oxy-combustionconditions are investigated using the model, and the parametersand properties use in the calculation are summarized in Table 5.

Fig. 21 shows the required time for drying a coal particle undertypical conditions in the mill and the furnace. The drying time ofa 100 mm coal particles is about 1 s in the primary gas stream, andabout 2.5 ms in the furnace, respectively, and it grows withincreasing particle diameter. An interesting observation is that thedrying times are similar in both air and CO2 environment. For thetypical size of a pulverized coal particle, the Reynolds number is

Table 5Parameters used in estimation of the heating and drying processes of single coal particle

Temperature (oC)

TN Tf Tboil

Drying in mill 105 e 100Drying in furnace 1000 1800 100Heating in furnace 1000, 900, 800 1800, 1700, 1600 e

small at the terminal velocity, such that the Nusselt number is closeto 2, and the convective heat transfer is similar to conductive heattransfer under stagnant condition. Therefore, the thermal conduc-tivity dominates the heat transfer rate. Given identical primarystream and combustion temperatures, the evaporation time is onlyslightly longer in themill, and shorter in the furnace under CO2-richenvironment than in air. The reason is that the thermal conduc-tivity of CO2 is slightly lower than air at the mill temperature, andslightly higher than air at the furnace temperature. However, thedifference is negligible if the primary stream temperature in themill, or surrounding gas and flame temperature in the furnace areidentical in the two environments. Similarly, Fig. 22 shows that theparticle heating rate under CO2-rich condition is only slightlyhigher than that in air.

The gas temperature and flame temperature are generally lowerin oxy-coal combustion with similar oxygen concentrations (seeSection 2.1.3). In order to investigate the impact of the potentiallylower temperatures on the evaporation time, additional calcula-tions were performed in which the temperatures in the CO2 gaswere reduced by 100 and 200 degrees from their original values,corresponding to TN ¼ 900 �C and Tf ¼ 1700 �C, and TN ¼ 800 �Cand Tf ¼ 1600 �C, respectively. The results of the new calculationare also shown in Fig. 22. The heating rate of a 100 mm coal particledecreases in the cases where the gas and flame temperatures are100 and 200 K lower than their original values. These resultsindicate that the particle heating rate is more significantly influ-enced by combustion temperatures than by gas compositions.

The analysis in this section indicate that employing a similarreheated recycled flue gas temperature will result in similar coalparticle drying performance to that of air-combustion, and the

.

Gas properties Coal properties

cpg ;m; r; k for pure substances arecited from [99] at Tg ¼ 1

2ðTN þ TboilÞ

cpg ¼ cH2O

kg ¼ 0:6kN þ 0:4kH2O [251]

ε ¼ 0:8,cp ¼ 1500 J=kgKhfg ¼ 2442 kJ=kgMoisture content ¼ 10%

Page 20: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Fig. 22. Predicted coal particle heating history in N2 and CO2 gas atmospheres. The gastemperature is set to 1000 �C and the flame temperature 1800 �C, respectively. The gasand flame temperature drops of 100 �C and 200 �C accounts for the possible lowertemperatures under oxyfuel condition.

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 175

evaporation time and heating rate does not change much if thecombustion temperature is maintained in oxy-fuel combustion. Itshould also be noted that this analysis is based on isolated singleparticle/droplet, the more likely phenomenon of group drying mayshow other dependencies, and should be studied in the future.

2.4. Coal devolatilization and char formation

During heating of the coal particles, volatile matter, includingtar, light gases, and pyrolysis water, are released into the gas phaseat a temperature above 300 �C. Devolatilization is an endothermicprocess and its kinetics and yields are strongly governed by thetemperature history (heating rate, peak temperature, and holdingtime), ambient gas compositions and pressure [114e121]. Devola-tilization kinetics and yields are usually investigated in inert gases,such as N2 and helium, to avoid the interactionwith ambient gases.Devolatilization in a hydrogen environment (a.k.a. hydro-gasification) is also of interest for the methanation process[116,122]. Similarly, the devolatilization characteristics in a highCO2 concentration environment, which is typical of oxy-fuelcombustion, are important and require more research.

Experimental studies on the volatile yields at different heatingrates using thermogravimetric analysis (TGA) and drop tubefurnaces (DTF) are summarized in Table 6. The volatile yields re-ported by different researchers vary considerably. Someresearchers observed higher volatile yields in CO2 at high temper-atures, and attributed this to the char-CO2 gasification reaction,while others observe similar volatile yields. For example, whencomparing the weight loss curves of a pulverized bituminous coalin N2 and CO2 atmospheres using TGA, Rathnam et al. [123]observed that coal devolatilization started at about 573 K and the

Table 6Lab/bench scale experiments on coal devolatilization in atmospheric N2 and CO2 environ

Author Facility Coal sample

Rathnam [22,123] TGA Coal A,B,C and DDrop tube furnace Coal A,B,C and D

Borrego et al. [125] Drop tube reactor High/Low volatile bituminousAl-Makhadmeh et al. [124] Entrained flow reactor Bituminous (KK) and lignite (LABrix et al. [126] Entrained flow reactor South American bituminous

weight loss history is similar for both atmospheres below 1030 K(see Fig. 23). However, the weight loss in CO2 is significantly higherwhen the temperature is higher than 1030 K due to the char-CO2gasification reaction. They also measured the volatile yields of fourcoals in a DTF with N2 and CO2 atmospheres at a nominal walltemperature of 1673 K. Once again, the volatile yields in CO2 areabout 5e25% higher than those in the N2 atmosphere.

Al-Makhadmeh et al. [124] have reached the same conclusion intheir pyrolysis experiments on a medium volatile bituminous coaland a lignite coal using an entrained flow reactor in CO2 and N2environments. Fig. 24 shows the mass release (or volatile yield) asa function of the wall temperature for both coals at 700e1150 �C.The volatile yield is almost identical in both environments at lowtemperatures, while higher mass releases are observed in CO2 attemperatures above 850 �C. At higher temperatures, the volatileyields are 10% and 11e14% higher in the CO2 environment than inthe N2 environment for bituminous coal and lignite coal, respec-tively. The higher deviation for the lower rank coal might be due toits relatively higher reactivity with CO2. More repeats of theexperimental results would gain a higher confidence. Measure-ments of the pyrolysis gas speciation of the lignite coal have alsobeen carried out at the end of the reactor in order to confirm thereason for the higher mass release in CO2. As shown in Fig. 25, COconcentration is indeed significantly higher, while H2 concentra-tion is considerably lower in a CO2 pyrolysis atmosphere. This maybe explained bywater-gas shift reaction occurring at the higher CO2partial pressure.

To the contrary, Borrego and Alvarez [125] conducted coalpyrolysis experiments using a drop tube reactor at 1300 �C, andreported a lower mass release in a CO2 pyrolysis environment forboth Colombian vitrinite-rich high volatile bituminous coal andWestern Canadian low volatile bituminous coal. The authors sug-gested that CO2 could be involved in the process of cross-linking atthe char surface of the resolidifying chars, which reduces theswelling of the particles and inhibits the volatile release.

Recently, Brix et al. [126] carried out pyrolysis experiments ona South American bituminous coal using an electrically heatedentrained flow reactor in N2 and CO2 environments. In contrast tothe observations reported in the previous paragraph, they found nodifference in char morphology, char N2-BET surface area, or volatileyields in the N2 and CO2 environments. The authors reviewed thediscrepancies from the previous experiments and suggested thatthe operating conditions used during these experiments, such asthe mixing between cold and hot gas streams, the particle heatingrate and the total particle residence times as summarized in Table 6,inevitably affected the results, These factors must be taken intoconsideration when the data are compared. They concluded thatthe char-CO2 gasification reaction does increase the volatile yieldduring devolatilization at high gas temperature and with longresidence times (0.62 s and 1 s) as reported by Rathnam et al. [123]and Al-Makhadmeh et al. [124], respectively; whereas the lowermass release in CO2 environment reported by Borrego et al. [125]might be due to the high amount of cold entrainment gas (w1/3)and short residence time (0.3 s).

ments.

Coal massflow rate

Size Carrier gasflow rate

Main gasflow rate

Nominaltemperature

Residencetime

(g/h) (mm) (L/h) (L/h) (�C) (s)

NA 63e90 NA NA <1200 24004e5 63e90 48 312 1400 0.6260 36e75 300 600 1300 0.3

) 500e2500 e 400 9200 700e1150 150 90e106 18% 82% 900e1400 0.15e0.297

Page 21: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Fig. 23. Normalized mass loss of a coal sample during pyrolysis in N2 and CO2

atmospheres in TGA, indicating an increased reactivity in CO2 at high temperatures,cited from [123].

Fig. 25. H2, CO and CH4 concentrations during a lignite coal pyrolysis in N2 and CO2

environments [124].

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214176

Correspondingly, the char morphology and specific surface areameasurements (BET) showed different trends in these abovementioned experiments. Larger surface areas were observed forcoals pyrolyzed under CO2 devolatilization conditions in [123] and[124] to different extents. This was attributed to the char-CO2gasification reaction and, as mentioned before, no significantdifferences were detected in the char sample in [126].

2.5. Ignition of coal particles

Ignition and combustion behaviors of pulverized coal particledepend on the coal rank, particle size, heating rate and oxygenconcentration in the environment. In general, two types of ignitionmechanism have been observed for pulverized coal combustion:homogeneous ignition and heterogeneous ignition [127e129].Homogeneous ignition is usually observed during high volatilebituminous coal combustion when volatile matter and O2 areheated to the mixture auto-ignition temperature. The high surfaceflux of volatile products forces the reaction zone away from thesolid surface, thereby preventing the solid material from oxygenattack [128]. On the other hand, sub-bituminous coal and lignite

Fig. 24. Mass release for a bituminous coal (KK) and a lignite coal (LA) during pyrolysisin both N2 and CO2 environments [124].

may experience heterogeneous ignition, in which ignition occursvia the oxidation of the coal surface [130], with or without frag-mentation [131]. In this section, the different ignition behaviors arediscussed in the presence of N2 or CO2 as diluent gas.

2.5.1. Theoretical analysis of homogenous ignition delayFor combustion of pulverized coal with high volatile contents,

homogenous ignition of the volatiles and oxidizer mixture occursfirst, followed by secondary ignition and burning of the charparticle. Therefore, it is instructive to investigate the factors gov-erning ignition delay of a mixture of gaseous volatiles and oxygenin a well-defined environment. Assuming that volatile combustionreaction can be simplified to a one-step global reaction rate inArrhenius form:

dcFdt

¼ cFAexpð�Ta=TÞ (16)

where cF is the molar concentration of the fuel and Ta is theactivation temperature. Following the derivation of Law [132], theignition delay can be related to the thermal properties of thereactant gas as follows:

si ¼cv�T20=Ta

�qcYF;0Aexpð � Ta=T0Þ

(17)

The Eq. (17) indicates that the homogeneous ignition delaydepends on the molar heat capacity of the mixture, cv, the reactionheat release, qc, the reactivity of the fuel, as well as other operatingconditions (the initial temperature, T0, and the initial mass fractionof fuel, YF;0). In the homogenous ignition of coal volatiles, ignitiondelay is influenced by the following:

� Temperature of the surrounding gas: Ignition delay of the vola-tiles is shorter at higher surrounding temperature, mainly dueto its exponential dependence on the temperature in theArrhenius reaction rate term [132].

� Heating and devolatilization rate: Unlike gaseous fuels, volatilesare released through coal devolatilization before homogeneousignition takes place. Therefore, the heating rate of coal particlesand correspondingly, the devolatilization rate, strongly influ-ence coal ignition delay.

� Oxygen concentrations: That ignition delay decreases withincreasing oxygen concentration has been observed in studies

Page 22: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 177

for both fine coal particles [133,134] and coal pellets [130] inO2/N2, and in an O2/CO2 environment in recent experimentalstudies [133,135,136] which will be discussed below.

� Coal volatile content: Similar to the trend of oxygen concen-tration, a higher volatile content also facilitates ignition. Forinstance, lignite and sub-bituminous coals have shorter igni-tion delays than bituminous coals because of their highervolatile yields during devolatilization [136,137].

� Molar heat capacity of the surrounding gas: Eq. (17) indicatesthat ignition delay is linearly proportional to the molar heatcapacity of the gas mixture. This is because the temperaturerise during ignition is inversely proportional to the heatcapacity of the surrounding gas, and a higher heat capacityslows down the temperature rise and chemical reactions. SinceCO2 has a molar heat capacity that is 1.66 times larger than thatof N2 (see Table 3), the ignition delay time is expected to belonger. Single coal particle experiments have also confirmedthis trend when replacing N2 with CO2, which will be discussedbelow in greater detail.

To examine these arguments qualitatively in a more idealizedenvironment, the ignition process of stoichiometric CH4-21% O2/79% N2 and CH4-21% O2/79% CO2 mixtures are investigated usingthe closed homogenous reactor model of CHEMKIN [138] with GRI-Mech 3.0 reaction mechanism [139]. In order to simplify theproblem, methane is used to be a substitute for the volatiles, and itis assumed to be well mixed with the surrounding gases in the coalparticle boundary layer and at atmospheric pressure. Anothersimplification is that reaction kinetics is much faster than heat lossto the reactor wall, and the reactor system is assumed to be adia-batic. Reaction kinetics and the energy equation are solved in thetransient problem, and the results show different trends for igni-tion delay in N2 and CO2 diluents. These assumptions may over- orunder- estimate the ignition delay. For instance, simplifying thevolatile composition and ignoring the presence of hydrogen mayoverestimate, while ignoring heat and mass transfer may under-estimate the ignition delay. Therefore, the results are onlyqualitative.

Fig. 26 shows the temperature profiles of the gas mixtures in theclosed homogenous reactor. Ignition delay is significantly longer inthe CO2 diluent gas than in the N2: the temperature rises take placeat 8 ms and 10.5 ms in N2 and in CO2, respectively. Moreover, the

Fig. 26. Comparison between the ignition delay of the CH4-O2-N2 and CH4eO2eCO2

mixture in a closed homogenous reactor. The symbol indicates O2/CO2 environ-ment, d indicates O2/N2 environment. Results show that the ignition delay is longer inCO2 diluent gas.

final combustion temperature is about 400 K higher in the N2diluent gas. It is worth noting that the initial conditions in thiscalculation are set to be close to that of Levendis’ experiments[136]. Although the time for particle heating and devolatilizationare not considered in this simulation, the results in Fig. 26 showsthe influence of the changing atmosphere on ignition delay quali-tatively. Simulation studies with more detailed consideration of theignition process are needed in order to develop a better under-standing of the ignition delay process in oxy-coal combustion.

2.5.2. Experimental studies on coal particle ignition delay in O2/N2

and O2/CO2

For the same coal type and operating conditions, ignition delayof single coal particles in both O2/N2 and O2/CO2 environments hasbeen experimentally studied. A series of experiments were con-ducted byMolina, Shaddix and coworkers in Sandia National Lab tostudy the ignition behavior of an individual coal particle [133,135]and of groups of particles [140] under O2/CO2 and O2/N2 condi-tions. The devolatilization time of Pittsburgh bituminous coal indifferent O2 concentrations in a laminar entrained flow reactorwere measured, using CCD and interference filter technology forimage capture and CH* chemiluminescence measurements. Theexperimental results on volatile ignition delay are shown in Fig. 27.The figure shows that the ignition delay time (si) is longer in an O2/CO2 environment than in an O2/N2 environment with similar bathgas temperature profiles along the reactor. It is also observed thatunder both conditions, the ignition delay time decreases withincreasing oxygen concentration. Shaddix and Molina [133,135]analyzed the trends of longer ignition delay in O2/CO2, and dis-cussed several possible effects at the presence of CO2, such asheating rate (heat transfer), homogeneous autoignition delay (heatcapacity), CO2 thermal dissociation (equilibrium), as well aschemical effect of CO2 on the radical formation. They suggested thatthe longer ignition delay in O2/CO2 mixtures is mainly due to thehigher heat capacity of CO2, its tendency to suppress radicalformation, or a combination of both [133].

More recently, Stivers and Levendis [136] measured the ignitiondelay times of four North American coals, including two lignitecoals, a sub-bituminous coal and a bituminous coal, in O2/N2 andO2/CO2 atmospheres. The coal particles were ground to a sizebetween 74 and 90 mm and were injected into a transparent quartztube in an electrically heated drop tube furnace. Gas stream of O2/N2 or O2/CO2 was introduced coaxially with the coal particles into

Fig. 27. Variation of ignition delay (si) for Pittsburgh bituminous coal for the differentgas mixtures, O2 concentration varies from 12%, 24%, to 36% [133]. Data are averagedfrom single particle images, with statistical error bars.

Page 23: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214178

the furnace, and the centerline furnace gases were measured toheat up and asymptotically reach a temperature 50 K below thefurnace wall set-point temperature of 1400 K, within a few centi-meters below the injector tip. As seen in Fig. 28, the experimentalresults showed longer coal particle ignition delay in the O2/CO2environment than in the O2/N2 environment with identical O2concentrations for all the coals tested. Moreover, the ignition delaytime tended to be shorter with increasing oxygen mole fraction inthe bulk gas under O2/CO2 conditions, but had no significant effectunder the O2/N2 conditions. The insignificant effect of oxygenconcentration under O2/N2 conditions may be because of the shortignition delay, and the differences are too small to be resolvedexperimentally. Further experiment measurements showed thatthe O2/CO2 mixtures were slower to heat-up than O2/N2 mixtures,therefore, the authors have attributed the longer ignition delay inO2/CO2 partly to the effect of volumetric heat capacity of the gasmixtures as described in 2.5.1, and partly to the slower heating rateof the background gas in the drop tube furnace which affects theconvective and conductive heat transfer to coal particles [141]. Notethat the longer ignition delay in O2/CO2 environment was observednot only for homogeneous ignitions of high volatile Pittsburghbituminous coal, but also in heterogeneous ignition circumstancesof lignite coals in the experiment of Stivers and Levendis [136].

The effect of CO2 on heterogeneous coal particle ignition has alsobeen investigated using wire-mesh reactor (WMR) by Qiao et al.[142]. The ignition temperatures of a Loy Yang brown coal anda Datong bituminous coal weremeasured in both O2/N2 and O2/CO2

environments with the objective of studying the ignition of coalparticles without the interference of volatiles burning. Since all thevolatiles produced by the coal particles are swept away by thecarrier gas flowing past the coal samples between the wire-mesh,the effects of volatile burning on the coal particle heterogeneousignition temperature are effectively eliminated. When air wasreplaced with a mixture of 21% O2/79% CO2, the average particleignition temperatures increased by 21 �C and 7 �C for the browncoal and bituminous coal, respectively. In order to confirm the

Fig. 28. Measured ignition delay times, si (ms), of four North American coals in O2/N2 andsingle particle images, with statistical error bar.

effect of the different diluent gases on the heterogeneous ignitiontemperature, the experiments were repeated using argon andhelium. As shown in Fig. 29, the experimental results show that theignition temperature with helium is significantly higher than withother gases, and ignition only occurs when increasing the samplemass to 4 mg. In general, the heterogeneous ignition temperatureincreases in the following order: Ar < N2 < CO2 << He. The authorsuggested that the significantly higher ignition temperature inhelium is because of its substantially higher thermal conductivity,which effectively cools down the char particle (see Table 7).Furthermore, a closer look at the experimental conditions showsthat the convective heat transfer coefficient of CO2 in the experi-ments is slightly higher than that of N2 due to its larger Reynoldsnumber, which may better explain the order of ignition tempera-tures in N2 and CO2 as observed in the experiment.

It should be noted that there are many differences between theexperiments conducted in the wire-mesh reactor by Qiao et al.[142] and the experiments conducted in the drop tubes by Levendiset al. [136] and Shaddix et al. [133,135]. In the drop tube experi-ments, coal particles were heated by the hot gas and the ignitionwas initiated via the homogenous/heterogeneous ignition of thevolatile matter and coal particles. In contrast, in the wire-meshreactor, the volatiles were not involved in the ignition of the coalparticles, and the particles were heated by conductive heat transferfrom the electrically heated wire-mesh, with the low temperaturecarrier gas acting as a cooling medium. In the latter case, heattransfer, i.e., the enhanced convective heat loss to the cooling gas,turns out to be one of the main reasons for the higher heteroge-neous ignition temperatures. Because of the distinctly differentboundary conditions in WMR, the assumption of an adiabaticexplosion model described in Section 2.5.1 does not hold. There-fore, the significantly higher volumetric heat capacity of N2 ascompared to Ar and He are not reflected by the coal particle ignitiontemperatures. It was also argued that the slightly higher ignitiontemperature in CO2 than N2 is because of the endothermic char-CO2gasification reaction [142].

O2/CO2 atmospheres with increasing O2 concentrations [136]. Data are averaged from

Page 24: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Fig. 29. Average particle ignition temperature for the Loy Yang brown coal during air(21% O2/79% N2) and the mixtures of 21% O2þ79% Ar, CO2, and He. The error barsindicate one standard deviation [142].

a b

Fig. 30. Variation of characteristic soot cloud size (a) and soot cloud combustiontemperature (b), during devolatilization of the Pittsburgh coal for different gascompositions, error bars represent 95% statistical error [133].

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 179

2.5.3. Burning of volatilesThe burning characteristics of coal vary significantly depending

on its rank, size and the operating conditions. For instance, Shaddixand Molina [133] found distinct volatile burning behaviors of twodifferent rank coals in their entrained flow reactor (EFR) experi-ments. During the ignition of a Pittsburgh high volatile bituminouscoal particle, a large hot volatile-burning soot cloud was observed,while lower radiation intensity was observed for the sootless BlackThunder subbituminous coal particle. Likewise, Levendis andcoworkers [136,143] also observed different volatile burningbehaviors in their drop tube furnace experiments: the combustionof bituminous coal was initiated by the ignition of released volatilematter which was sooty and bright, whereas lignite coals experi-enced extensive fragmentation during pyrolysis and combustionwithout clearly distinguishable volatile and char combustionstages. The different volatile burning phenomena are related to coalrank because soot is produced from the burning of long hydro-carbon chains in bituminous coal tar, while tar yields are muchlower in lignite and anthracite devolatilization. Since it is difficult tomeasure the volatile burning behavior of lignite and subbituminouscoals, the volatile burning characteristics were only reported forbituminous coals.

Lower combustion temperatures and longer burning times havebeen reported during the combustion of bituminous coal volatilesunder CO2-rich conditions. Shaddix and Molina [133,135] investi-gated the combustion of Pittsburgh coal particles. As shown inFig. 30, when replacing N2 with CO2, there appeared to be a trendtoward a slightly larger soot cloud size and a lower soot cloudtemperature. The same trends are confirmed by Bejarano and Lev-endis [143]: Fig. 31 shows that themeasuredburning temperature ofthe bituminous coal volatiles was lower, and burnout time waslonger in O2/CO2 environments over an oxygen mole fraction rangeof 20%e80%. Moreover, based on the higher volatile burningtemperature and the lower particle temperature, it is believed thatthe volatiles burn in a diffusionflame surrounding the devolatilizing

Table 7Comparison of physical properties of He, Ar, N2 and CO2 at 1 atm and 700 K. Data are cit

Property Unit He

Thermal conductivity k W=mK 0.2811Volumetric heat capacity rCp kJ=m3K 0.3618Thermal diffusivity a m2=s 7.768e-4

coal particle [133]. Lower burning temperatures and longer burnouttimes in the CO2 diluted environment are likely to be caused by thelower volatiles and/oroxygendiffusion rate (similar to the argumentof the lower char combustion rate, which will be discussed in detailin Section 2.6.2).

2.6. Oxy-char combustion

The effects of CO2 on char combustion/gasification have beenextensively discussed for air-fired coal combustion [144e147], andmore recently, for oxy-fired coal combustion [123,143,148149]. CO2is not an inert diluent in combustion; thus, it may influence charcombustion via several possible mechanisms: (a) the reducedoxygen mass transfer in CO2, (b) the lower temperature due to thehigher heat capacity of CO2, and (c) the char-CO2 gasificationreaction [148].

Many experimental studies have been performed in recent yearsto investigate the effect of a CO2-rich environment on thecombustion rate of coal char. Care should be taken when inter-preting these studies because experimental results are stronglydependent on the operating conditions. For instance, char oxida-tion may take place in different regimes, either kinetic-controlled(Zone I), diffusion-controlled (Zone III), or controlled by both(Zone II), depending on the experimental temperature range, thecoal rank and particle size used in the experiments, as well as theoxygen level in the ambient gas, among other factors. As a result,the char conversion rate under oxy-fuel conditions may becomehigher or lower than that in air-fuel mixtures.

With careful investigation of the experimental conditions usedin the studies, the results can be classified into three regions, asillustrated in Fig. 32, based on the oxygen concentrations andnominal temperatures in these experiments. Note that the nominaltemperatures of the WMR experiments are the measured particletemperature; while in TGA, DTF and EFR experiments, because thechar particle temperatures are not available in some of the studies,

ed from [99].

Ar N2 CO2

0.03396 0.04961 0.049300.3619 0.5353 0.86349.3832e-5 9.2666e-5 5.7092e-5

Page 25: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

a b

Fig. 31. Average temperature (a) and burnout time (b) For bituminous coal volatilesburning in O2/N2 and O2/CO2 at Tfurnace ¼ 1400 K, error bars show one standardexperimental deviation [143]. Broken lines — correspond to burning in oxygen-nitrogen mixtures while continuous lines correspond to burning in oxygen-carbondioxide mixtures.

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214180

the nominal temperatures are the gas temperatures, which arehundreds degree lower than the burning char particle temperature,depending on the oxygen mole fraction and coal rank. For instance,the temperature difference between the char particle and the gas is100e600 K in the Sandia EFR with an oxygen mole fractionbetween 12% and 36% [148], and is 400e1600 K in the NortheasternUniversity DTF with an oxygen mole fraction between 20% and100% [143].

� Region A- Low temperatures at any oxygen concentration: Thecharoxidation reactiondominates the char consumption,whichis kinetically controlled at these low operating temperatures.Most TGA experiments fall in this region, and the experimentalresults show similar reaction rates in O2/N2 and O2/CO2 envi-ronments. Clearly, when the oxygen concentration approaches100%, the difference between the two environments becomes

Fig. 32. Char oxidation/gasification experiments in oxy-fuel conditions. The diagramshows three regions where the experiments were conducted. A: At low temperatures,reactions rates are the same in both O2/N2 and O2/CO2 conditions; B: At high oxygenlevel and high temperatures, reaction rates are lower in oxy-fuel conditions; and C: Atlow oxygen level and high temperatures, reaction rates are higher in oxy-fuel condi-tions. The error bars show the range of operating conditions, colors show unchanged(black), decreased (blue), or increased (red) char consumption rate. (For interpretationof the references to colour in this figure legend, the reader is referred to the webversion of this article.)

negligible and the char consumption rate is the same even athigh temperatures, as shown in the small narrow region in thetop right of Fig. 32.

� Region B- High oxygen concentrations, high temperatures: Thechar oxidation reaction dominates the char consumption, butchar consumption is either in Zone II (internal diffusioncontrolled) or Zone III (external diffusion controlled). DTF andEFR experiments are typically conducted in this region, and theresults show lower char consumption rates in O2/CO2 condi-tions because of the lower oxygen diffusion rate in CO2.

� Region C- Low oxygen concentrations, high temperatures: Chargasification reactions become significant at high temperaturesand low oxygen concentrations. Experiments and numericalstudies have shown higher char consumption rates due to thegasification reactions. Similar to the exemption in Region A, thegasification reactions may also become dominant at lowtemperatureswhen the oxygen concentration approaches zero,as shown in the bottom left of Fig. 32.

The partitioning in Fig. 32 is only qualitative and the boundariesof different regions may differ due to the differences in apparentchar combustion behavior and reactor architecture. The dominantmechanisms in Regions A, B and Cwill be discussed in greater detailin the following sub-sections.

2.6.1. Kinetics of char oxidation in oxy-fuel combustionThe kinetics of the char oxidation reactions are usually

measured at relatively low temperatures (conditions in Region A) toensure that the reaction is kinetically controlled. TGA experimentshave been used to measure the char consumption rate in CO2-richenvironments, and the results have shown that the intrinsic charoxidation kinetics is not affected. For example, Varhegyi and co-workers [150,151] conducted TGA experiment in O2/CO2 and O2/Ar mixtures at atmospheric and elevated pressures and did notobserve any specific effect when replacing Ar by CO2 under theirexperimental conditions (5e100% O2, 400-900 �C, 1 bar and 6 bar),except that the thermal decomposition of a mineral (calciumcarbonate) component shifted to a much higher temperature in theO2/CO2 bath gas.

Liu observed the same trend in TGA experiments with a high-volatile U.K. bituminous coal and an anthracite coal in non-isothermal heating processes (2.5e12.5 K/min) [152]. Fig. 33compares the normalized reaction rates with increasing charconversion at a heating rate of 2.5 K/s in 10% O2/90% CO2 and 10%

a b

Fig. 33. Comparison of the reaction rates versus conversion between the combustionin 10% O2/90% CO2 and the combustion in 10% O2/90% N2 (heating rate ¼ 2.5 K/min):(a) U.K. Anthracite char and (b) U.K. Bituminous char [152].

Page 26: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 181

O2/90% N2 environment. Results show that replacing the inertnitrogen gas with CO2 has very little influence on the measuredcombustion rates of coal chars at all char conversion levels. Thesame conclusion was reached in TGA measurements in thetemperature range of 723e1163 K for anthracite and bituminouscoals [153]. The negligible effect of carbon dioxide on the charreaction rate was attributed to the much lower reaction rate of thechar-CO2 reaction than that of the char-O2 at the temperature andreactant gas composition in their experiments.

To verify this explanation, Varhegyi et al. [150] measured thechar consumption rate in pure CO2, and found that the char-CO2gasification reaction rate at 900 �C has the same magnitude as thatin 5e29% O2 around 450 and 500 �C. That the intrinsic kinetics ofthe char-O2 reaction is about three to four orders of magnitudelarger than those of the char-H2O and char-CO2 reactions, whereasthe char-H2O reaction is several times faster than the char-CO2reaction, have been reported previously [144,146,147]. Consideringthe substantially different intrinsic rates between the oxidation andgasification reactions, the gasification reactions only affect the charconsumption rate in O2-deficient environments, which will bediscussed in 2.6.3.

2.6.2. Effect of diffusivityWhen approaching higher combustion temperatures while

maintaining a high oxygen concentration in oxy-fuel combustion,the operating conditions transit from region A to region B in Fig. 32,in which DTF and EFR experiments are typically conducted. In thisregion, the char-O2 oxidation reaction still dominates, similar to thecase in region A. However, char oxidation (char þ O2) becomesprogressively more diffusion controlled with increasing charparticle temperatures, and for larger particle sizes. Therefore,oxygen diffusion turns out to be the controlling process in deter-mining the char consumption rate in this region.

Measurements of temperatures and burnout times of single charparticles have been conducted in both N2 and CO2 diluent gases.Bejarano and Levendis [143] investigated the burning of bitumi-nous coal, lignite coal, and a spherical synthetic char at increasingoxygen concentrations in both N2 and CO2 diluent gases. Theyconducted particle temperature and burnout time measurementsin a 4.2 kWelectrically heated laminar flow DTF using a three-colorpyrometer at relatively high oxygen concentrations (20%e100%)and high temperatures (1400 K and 1600 K). The experimentalresults showed that the coal particle temperatures are lower, andthe burnout times are longer, in O2/CO2 than in O2/N2. For example,when operating the furnace wall temperature at 1400 K, thedifferences in char temperature between the two conditions, forbituminous coal and lignite coal, were on the order of 200 K and120 K, respectively. Similarly, Shaddix and Molina [133,148]measured the char particle temperature of a Pittsburgh high-volatile bituminous coal and a Powder River Basin subbituminouscoal in Sandia’s EFR at a gas temperature of 1700 K over an oxygenconcentration range of 12e36% in N2 and CO2 diluent gases. Lowerchar particle temperatures and burning rates were observed forboth coals in the O2/CO2 environment. After accounting for thediffusion limitation of oxygen through the particle’s boundarylayer, the deduced intrinsic kinetics of the char oxidation reactionswere found to be the same for O2/N2 and O2/CO2 conditions.

The carbon conversions of sampled char particles during oxy-char combustion in entrained flow reactors also show the sametrend. Brix et al. [126] measured the char conversion profiles in anEFR under O2/N2 and O2/CO2 conditions with a coal feeding rate of50 g/h. At the lower temperature range of 1173e1373 K and with anoxygen concentration of about 6%e28%, they observed no signifi-cant difference in the char conversion rate between the two envi-ronments, which confirms the trends in region A. When increasing

the wall temperatures to 1673 K, where char oxidation is controlledby diffusion, the char conversion rates in CO2 were lower. Likewise,Al-Makhadmeh et al. [124] studied char combustion in a once-through 20 kW EFR at 1300 �C with an oxygen concentration of5%e15% under O2/N2 and O2/CO2 conditions. The chars used for thecombustion studies were produced in N2 and CO2 devolatilizationenvironment, respectively. Similarly, a comparison between thetwo conditions shows that the char conversion and oxygenconsumption are indeed slower in the CO2 balance gas. Moreover,they found that the difference between the two combustion envi-ronments is less significant at lower oxygen concentrations, whichindicates that the diffusion effect becomes less dominant when theoxygen levels decrease (approaching region C).

In this paper, a simplified model for char particle combustion isused to investigate the impact of mass transfer on char oxidation inboth air-fired and oxy-fired environments. Studies on carbon orcoal char combustion are voluminous, and various models, such asthe Single Film Model (SFM) and Double Film Model (DFM), havebeen proposed to describe the mass and heat transfer in theboundary layer surrounding a combusting char particle. Thedifference between these models is mainly the approximation ofCO burning locations [154]. Mitchell et al. [155] modeled the COoxidation process in coal boundary layers and suggested that littleCO2 is formed in the boundary layer when the char diameter issmaller than 100 mm, which warrants the applicability of SFM forthe size range of pulverized coal combustion. Therefore, SFM isapplied in this study to examine the fundamentals of single charparticle combustion in O2/N2 and O2/CO2, and the experimentalresults in [143] are used to verify the calculations from this model.

The char particle is described as a shrinking homogenous sphereparticle surrounded by a chemically frozen boundary layer [156],and the burning process is considered to be quasi-steady ina stagnant gas environment (O2/N2 or O2/CO2 with increasing O2mole fractions). Oxygen diffuses toward the char particle throughthe dilution gas (either N2 or CO2) and reacts at a finite ratewith thechar surface to produce CO and CO2. Based on their experimentalfindings, Tognotti et al. [157] suggested that the CO2/CO ratioformed by the heterogeneous reaction on the char surface is lessthan 0.1 at normal combustion temperatures. Therefore, thesimulation only considers CO as the combustion product, whichdiffuses away from the char surface. Diffusion and reactions insidethe char particle are neglected.

Following the derivation in [109], the reaction heat release fromchar surface is balanced by thermal conductive and radiative heattransfer to the surrounding gas:

_mCDh ¼ �kg4pr2sdTdr

rs þ εs4pr2s s�T4s � T4sur

�(18)

where the temperature gradient at the particle surface is given by:

dTdr

jrs

¼ _mCcpg4pkgr2s

26664ðTN � TsÞexp

�� _mCcpg4pkgrs

1� exp�� _mCcpg

4pkgrs

�37775 (19)

The O2 mass transfer in the particle boundary layer followsFick’s Law:

_mC ¼ 4pr2rD�1þ YO2

=nO2

� d�YO2=nO2

�dr

(20)

where nO2is the oxygen to fuel mass ratio. Char consumption rate is

estimated using a first order surface reaction:

_mC ¼ 4pr2s kCPO2;s (21)

Page 27: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Fig. 35. Temperature profiles in the boundary layer around a 50 um diameter burningchar particle with radial coordinate, predicted using the Single Film Model in air-firedand 21% O2/CO2 conditions.

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214182

where the partial pressure of oxygen at the particle surface is

PO2;s ¼ PYO2 ;sMWmixMWO2

, and the kinetics of char-O2 surface reaction

rate, kC , is taken from the experimental results of Field [158]:

kC ¼ 0:8596 exp�17967

Ts

�(22)

Note that the pre-exponential factor in kC is different from theoriginal reference due to unit conversion. The diffusivities used inthemass diffusion equations are for the binary diffusion of O2 in theN2 or CO2 bath gases (see Table 3).

The differential Eqs. (18), (20), and (21) with appropriateboundary conditions are solved simultaneously to find the charconsumption rate, _mC , char surface temperature, Ts, and the charsurface oxygen mass fraction, YO2 ;s for a fixed radius, rs, in quasi-steady state.

Fig. 34 to Fig. 37 shows the experimental and simulation resultsusing the SFM model, from which the effect of diffusivity in N2/O2

and O2/CO2 combustion can be seen. Fig. 34 illustrates the simu-lation results for species mole fraction profiles under both envi-ronments with an O2 concentration of 21%. Oxygen diffuses towardthe particle in the bath gas (N2 or CO2), as shown in the figure.Recalling that the O2 diffusivity in CO2 is 0.78 times that in N2 (seeTable 3), the oxygen diffusion flux, and thus the carbon consump-tion rate is correspondingly lower in the CO2-rich environment.Fig. 35 shows the predicted temperature profile in the boundarylayer of the particle. Lower temperatures, especially at the particlesurface (R/rp¼ 1), is observed under oxy-fuel combustion conditionwhen compared with the temperatures in air combustion. The charsurface temperature in 21% O2/CO2 is about 190 K lower than that inair-fired combustion. The temperature profile results froma balance between the reaction heat release and the conductive andradiation heat transfer at the particle surface during char burningprocess, and it reflects the extent of the char burning intensity. Alower O2 diffusion rate in oxy-fuel combustion translates intoa lower char consumption and heat release rate. Therefore, theparticle surface temperatures are lower than in air combustion.

Fig. 36 and Fig. 37 show the measured and predicted particlesurface temperatures and burnout times, respectively, under O2/N2and O2/CO2 conditions with increasing O2 mole fraction. Both theexperimental results from Bejarano and Levendis [143] and the

Fig. 34. Species (O2, N2, CO, and CO2) mole fraction profiles in the boundary layeraround a 50 um burning char particle in the outward radial direction, predicted usingthe Single Film Model in air-fired and 21% O2/CO2 conditions.

modeling results show lower particle temperatures and longerburnout times under O2/CO2 conditions. Similar trends have beenreported by Shaddix and Molina [148] in their entrained flowreactor (EFR) experiments: the combusting char temperature islower in CO2 than in N2 at the same gas temperature (1500e1750 K)and O2 concentrations (12%, 24% and 36%). The predicted particlesurface temperatures under O2/N2 conditions match well with theexperimental data. However, the predictions overestimate theparticle surface temperatures by about 200e400 K under O2enriched conditions (oxygen mole fraction higher than 0.8). Thisdiscrepancy may be because the simulation did not consider thedissociation reactions and equilibrium at very high combustiontemperatures. Moreover, the assumptions such as CO as the soleheterogeneous reaction product and ignoring its oxidation in theboundary layer may not be valid at high particle temperatures andhigh oxygen mole fractions [155,157]. The gasification reactionsmay also become important at high char temperatures [149,159]. Amore comprehensive consideration of the physics and chemistryfundamentals in O2/CO2 conditions is needed.

Fig. 36. Particle surface temperature of a 50 um char particle in O2/N2 and O2/CO2

mixtures at Tfurnace ¼ 1400 K. Lines show the predicted results using the Single FilmModel, and markers show the experimental data in [143]. Continuous lines d andbroken lines correspond to O2/N2 and O2/CO2 conditions, respectively.

Page 28: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Fig. 37. Burnout times of a 50 um char particle in O2/N2 and O2/CO2 mixtures atTfurnace ¼ 1400 K. Lines show the predicted results using the Single Film Model, andmarkers show the experimental data in [143]. Continuous lines d and broken linescorrespond to O2/N2 and O2/CO2 conditions, respectively.

Fig. 38. Arrhenius plot of NL coal char oxidation/gasification reactions rate. The trendlines are based on experimental data from literature [147]. Zone I, II and III indicatekinetic controlled, internal diffusion controlled and external diffusion controlledheterogeneous reactions, respectively.

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 183

2.6.3. Effect of gasification reactionsThe effect of the gasification reactions on char consumption in

oxy-char combustion has attracted much interest in recent years.Char gasification reactions are different from char oxidationreactions:

� Gasification reactions are highly endothermic and thus requireheat transfer to the char particle. The reaction energies are172 kJ/mole-carbon for char-CO2, and 131 kJ/mole-carbon forthe charesteam reaction [149]. In coal gasification processes,the external energy required for the gasification reactions isusually provided from the heat released by oxidation reactions.

� Gasification reactions have much higher apparent activationenergies. The activation energies of the char-CO2 and char-H2Oreactions are 230e270 kJ/mol and 190e270 kJ/mol, respec-tively, while the char-O2 reaction activation energy is about160 kJ/mol [149]. Therefore, gasification reactions only occur athigh temperatures.

� Gasification reactions are kinetically controlled under mostcombustion conditions due to the slow reaction rates.

Fig. 38 shows an Arrhenius plot of NL coal char oxidation andgasification reaction rates obtained from pressurized drop tubefurnace (PDTF) and TGA, based on the experimental work of Kaji-tani et al. [147]. As has been discussed in Section 2.6.1, the order ofmagnitudes of char-CO2 and char-H2O reaction kinetics are roughly4 times lower than that of char-O2 at typical burning temperatures.Therefore, as illustrated in region C of Fig. 32, gasification reactionscan only contribute to a higher char consumption rate at hightemperatures with low oxygen concentrations in CO2-rich envi-ronments. As long as the oxygen concentration is high, thecontribution of gasification reactions is relatively smaller than thatof oxidation reactions, and the char consumption rate is lower inCO2-rich environments than in N2-rich environments, as would beexpected in region B.

The growing effect of gasification reactions at high temperaturesand lower O2 concentrations has been observed in air-fired coalcombustion modeling studies. Mann and Kent [145] numericallystudied the heterogeneous char reactions and found that theunburned carbon (as the percentage of the dry ash free coalentering the furnace) is reduced from 3.4% to 0.7% due to char-H2O

and char-CO2 gasification reactions in air-combustion. Stanmoreand Visona [146] further demonstrated the importance of the char-H2O reaction during char combustion in Zone II (internal diffusioncontrolled) and claimed that the contribution of gasification reac-tions was greater than estimated by Mann and Kent.

Considering that CO2 partial pressure in oxy-coal combustion ismuch higher than in the air-fired case, gasification should becomemore prominent at the late stages of oxy-fuel combustionprocesses. Rathnam et al. [123] compared the burnout of four coals(coal A, B, C and D) at oxygen concentrations of about 3e21% ina DTF at 1673 K, and the results are shown in Fig. 39. They observedthat two coals (A and B) have higher once-through conversions inthe O2/CO2 environment. The authors attributed the higherconversions to the char-CO2 gasification reaction. To verify thiseffect, the reactivity of coal-A char formed in N2 at 1673 K in the DTFwas measured at different O2 levels with N2 and CO2 balance gases.It can be observed in the TGA experiment shown in Fig. 40 that thechar reaction rate only becomes noticeably higher at temperaturesabove 760 �C with a very low O2 concentration of 2%. Likewise, thechar-H2O gasification reaction’s effect on char consumption wasalso noticed in the char yield of Loy Yang brown coal at a very lowoxygen concentration and high temperature of 800 �C, in theexperiments conducted by Guo et al. [160] using a wire-meshreactor. They observed that the char consumption rate ina mixture of 2000 ppm oxygen and 15% steam with Ar diluent gaswas higher than in 2000 ppm oxygen with Ar diluent gas.

It is worth noting that the previously discussed observations inRathnam et al.’s experiments (across region B and C of Fig. 32) [123]and in Brix et al.’s experiments (in region B of Fig. 32, also seeSection 2.6.2) [126] seem contradictory, although both experimentswere conducted in a suspension fired condition at a high temper-ature (1673 K) and low oxygen concentrations (around 3%e6%).This may be because of the different residence times, or thedifferent definitions of nominal oxygen concentration in theirexperiments. Brix et al. [126] used the measured local oxygenconcentration after volatile consumption and before char oxidation,whereas in Rathnam et al.’s experiments [123], the measurementlocation of the oxygen concentration and stoichiometry are notstated [123]. If the measurements of oxygen concentration weredone before volatile combustion, the actual oxygen concentrationsfor char combustion may be lower than the reported values, and

Page 29: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

a b

c d

Fig. 39. Comparison of burnout of coals at various O2 levels in air and oxy-fuel atmospheres in DTF at 1673 K, with error bars representing the standard deviation obtained frommultiple samples at each experimental condition [123].

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214184

therefore, the contribution of the gasification reactions becomesmore significant.

A simplified multi-surface reaction model is used to estimatethe contribution of gasification reactions to char consumption inair-fuel and oxy-fuel combustion. The char is assumed to be a 50-mm diameter solid sphere with heterogeneous reactions takingplace on its external surface. The reactions are simplified to be 1storder following Smith’s approach [161], but three global charoxidation and gasification reactions are considered:

Reaction ðIÞ: 2CðsÞþO2/2CO (23)

Reaction ðIIÞ: CðsÞþH2O/COþ H2 (24)

Reaction ðIIIÞ: CðsÞþCO2/2CO (25)

Fig. 40. Reactivity comparison of Coal A char, prepared in N2 in DTF at 1673 K, in airand oxy-fuel atmospheres in TGA indicating a higher char reaction rate for the 2% O2/CO2 case at higher temperatures [123].

The overall char consumption rate is the summation of theseindividual surface reaction rates:

_mC ¼Xni¼1

_mC;i (26)

The char reaction rate for reaction i ( _mc;i) depends on the rate ofreactant transport by external diffusion and the surface reactionkinetics

_mC;i ¼ 4pr2pPi;gkdiff ;ikkin;i

kdiff ;i þ kkin;i(27)

where kinetic rate constants of the char surface reactions are inArrhenius form:

kkin;i ¼ AiTbip exp

�EiRuTp

�(28)

The kinetics parameters for external surface reactions that areused in this estimation are taken from references [158,162,163], seeTable 8. The orders of particle temperature, bi, are all zeros for thesereactions.

Diffusion often controls the char-O2 reaction rate at hightemperatures. Following the approach described in [161], the cor-responding mass transfer limited reaction rate constantkdiff ;i, canbe expressed as:

kdiff ;i ¼ Ci

��Tp þ TN

� 2�0:75

dp(29)

where Ci is the overall mass diffusion-limited constant:

Ci ¼ niMWc

MWi$

MWRuT1:750

$Sh$Di;0

�P0P

�(30)

where ni is the stoichiometric coefficient of carbon in reaction I,II and III (Eqs. (23)e(25)), MW is the average molecular weight ofthe gas mixture in boundary layer, the various Di;0 are taken from

Page 30: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Table 8Char surface reaction parameters.

Reaction Ai Ei Ref Ci

kg=ðm2sPaÞ kJ=mol s=K0:75

I 0.005 74 [158, 162] 4.13E-12a, 5.32E-12b

II 0.00192 147 [162, 163] 4.12E-12a, 5.77E-12b

III 0.00635 162 [162, 163] 1.69E-12a, 1.72E-12b

a for oxy-fired condition.b for air-fired condition.

Fig. 41. The contribution of gasification reactions to the overall carbon consumption ofa 50 mm char particle as function of particle temperatures. The estimations are per-formed under O2 rich and deficient conditions in air-fired and oxy-fired coalcombustion, as shown in Table 9.

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 185

[164]. Note that Ci is different for air-combustion and oxy-combustion because of the different diffusivities in the N2-rich orCO2-rich environment.

The contribution of the gasification reactions to the total carbonconsumption was estimated under O2-rich and O2-deficientconditions, in air-fuel and oxy-fuel combustion, as shown inTable 9. Fig. 41 shows the contribution of gasification reactions tothe overall carbon consumption with increasing particle surfacetemperatures. At the same particle burning temperatures, thecontribution of the gasification reactions is increased in oxy-fuelcombustion because of the higher CO2 partial pressure. Moreover,the gasification reactions become more significant in O2-deficientconditions. For example, the contribution of gasification reactionsat 2000 K increases to approximately 40% for the oxy-fuel O2-deficient condition, compared with w10% for the air-fired O2-deficient condition.

Recall that since the char-O2 reaction becomes slower due to thelower O2 diffusion in CO2, as discussed in 2.6.2, the effects of havingCO2 as a bath gas are two-fold. In the early stages of charconsumption in oxy-fuel combustion, the char-O2 oxidation reac-tion dominates and the reaction rate is lowered due to the effect ofdiffusion inhibition. However, in the late stages of char consump-tion, the carbon consumption rate increases as the char-CO2 andchar-H2O gasification reactions become significant. Note that sincethe understandings on char gasification kinetics and its interactionswith oxidation reactions are still limited, the estimate above shouldbe only considered as qualitative and coal-type dependent.

2.6.4. Effect of heat capacityTo identify the impact of the higher molar heat capacity of CO2

on the reduced surface temperature and consumption rate duringoxy-char combustion, Shaddix and Molina [148] simulated thesteady-state char burning process using a detailed computationalprogram SKIPPY (Surface Kinetics in Porous Particles, developed atthe University of Sydney) in which the mass and energy conser-vation, the multi-component diffusion, and the reaction within thechar pores are considered. When all the thermal properties of CO2except the diffusivity are artificially assumed to be identical tothose of N2, the temperature profiles under oxy-fuel conditions donot differ significantly from the full simulation. This indicates thatthe thermal properties have only a minor impact on the surfacetemperature and consumption rate for the combustion of a singlechar particle.

However, it is noteworthy that the conclusion above is madewith the assumption of fixed gas stream temperature in a dilutedcoal burning environment in the EFR. In practice, the higher heat

Table 9Gas compositions used for the carbon consumption calculation.

Conditions O2 H2O CO2 N2

% % % %

Air-fired O2 rich 21 0 0 79Air-fired O2 deficient 3.3 6 14.7 76Oxy-fired O2 rich 21 0 79 0Oxy-fired O2 deficient 3.3 6 90.7 0

capacity of CO2 does reduce the combustion temperature in oxy-fuel combustion (see Section 2.1.3). Without a fixed bulk gastemperature, the lower surrounding gas temperature will result ina lower particle surface temperature, and hence a lower burningrate [165].

2.6.5. Effect of pressureStudies of the effect of pressure on char oxidation in oxy-fuel

combustion are still scarce because the focus so far has been onatmospheric oxy-fuel systems. However, qualitative trendsregarding the impacts of pressure on char oxidation and gasifica-tion can be obtained using observations made in previous coalgasification studies, such as that of Niksa et al. [165] and Liu et al.[166] which reviewed char oxidation and gasification characteris-tics under elevated pressures from selected coal reaction databases.

The char-O2 reaction rate rises with increasing the oxidizerconcentrations at a constant total pressure. Experiments conductedin a drop-tube reactor [167] showed significantly higher burnoutwith increasing O2 levels for Westerholt coal at a total pressure of1.5 MPa and a reactor temperature of 1200 �C. This is becauseincreasing the O2 level at a constant total pressure raises theoxidizer bulk diffusion flux. On the other hand, the consumptionrate does not change much with increasing the total pressure atconstant O2 mole fraction under diffusion controlled conditions.Fig. 42 shows the burnout history of a high volatile bituminous coalat higher total pressures (0.1e1.5 MPa) and at a constant O2 molefraction of 21%. The reported degree of burnout increases onlyslightly while the total pressure increases from atmospheric pres-sure to 1.5 MPa. This is because the diffusion rate is independent oftotal pressure at a constant O2 mole fraction: the effect of the rise inthe bulk partial pressure is canceled by the drop in the massdiffusion-limited constant (also see Eq. (30)).

The pressure dependence of the char-H2O and char-CO2 reactionis more straightforward. As shown in Fig. 38, gasification arekinetically controlled up to about 1200e1400 �C, therefore theirreaction rates increase with increasing partial pressure of H2O andCO2 at moderately high temperature during coal combustion andgasification. Fig. 43 shows the effect of pressure on the gasificationreactions at 1000e1300 �C. Since the partial pressure is the productof the total pressure and the mole fraction, the reaction ratesincrease by either raising the total pressure at a constant molefraction (in Fig. 43(a)), or increasing the mole fraction at a constant

Page 31: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Fig. 42. Burnout history of a high volatile bituminous coal in an EFR at 21% O2 and800 �C. Coal particle sizes are from 315 to 500 mm, cited from [165].

a

b

c

Fig. 43. Pressure effects on char gasification reactions. (a) Char conversion history ofa bituminous char in steam and CO2 at elevated pressures in wire-mesh reactor(WMR). (b) Reaction rates of abituminous char at increasing reagent mole fractionsunder a constant total pressure. (c) Reaction rates of a bituminous char at a constantpartial pressure and increasing total pressure [166].

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214186

total pressure (in Fig. 43(b)). Conversely, maintaining the samepartial pressure of H2O or CO2 while raising the total pressure doesnot affect the reaction rates much (in Fig. 43(c)) [166]. Diffusioneffects will become important at high temperatures, above about1400 �C [147].

Comparing Fig. 42 and Fig. 43(a), the effects of the total pressureat a constant reagent mole fraction are distinctly different for charoxidation and gasification reactions, although the char conversionrates are much slower for the gasification reactions. The impacts ofpressure on the char oxidation and gasification reactions aredifferent because of the nature of their reaction regimes. As dis-cussed in Section 2.6.3, for pulverized coal combustion conditions,char oxidation reactions are in the bulk diffusion controlled regime(Zone III) or internal diffusion controlled regime (Zone II), whilechar gasification reactions fall in the kinetically controlled regime(Zone I) or the internal diffusion controlled regime (Zone II),depending on the char particle size, reactivity and operatingconditions. Therefore increasing the total pressure alone would notincrease the coal particle-burning rate significantly in the near-burner zone in oxy-fuel combustion. It is more effective toincrease the O2 level by adjusting the FGR in the burner stream oroxygen injection. Moreover, raising the total pressure may facilitatethe gasification reactions and thus increase the char conversion inthe late stages of oxy-fuel combustion.

2.6.6. SummarySection 2.6 provides a review of the characteristics of oxy-fuel

char combustion in a CO2-rich environment. In pulverized coalcombustion, coal particles are burned mostly in Region B andRegion C, where the effects of the physical properties and gasifi-cation reactions on char conversion rates are important. The mainconclusions in this section are:

� TGA Experimental results for the char oxidation at lowtemperatures (as shown in Zone I of Fig. 38) do not show anydifferences between O2/N2 and O2/CO2 environments, whichindicates that the intrinsic kinetics of the char-O2 reaction arethe same in the O2/CO2 or O2/N2 environments.

� The char-O2 reaction rate becomes lower in a CO2 diluent underdiffusion controlled conditions (as shown in Zone II and III ofFig. 38) due to the lower O2 diffusivity in CO2. Experimentalmeasurements and predictions at high temperatures with highoxygen concentrations have confirmed this trend.

� The higher molar heat capacity of CO2 lowers the adiabaticflame temperature in O2/CO2 condition (with similar oxygenmole fractions), and hence the reaction rates. However, givenidentical surrounding gas temperatures (by decreasing flue gas

Page 32: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 187

recycle ratio), higher heat capacity alone does not result inlower char temperature or oxidation rate.

� Since gasification reactions are much slower than the char-O2reaction, they are kinetically controlled at typical combustiontemperatures. The contribution of gasification reactionsbecomes significant in the late stages of oxy-fuel combustionwhere the O2 concentration is low and the surrounding gastemperature is high.

� Raising the total pressure while maintaining the oxygen molefraction in the oxidizer stream does not increase the charoxidation reactions; however, doing so increases the gasifica-tion reaction rates under oxy-fuel conditions.

It should be noted that the above conclusions are based onanalysis using simplified models. Other mechanisms might alsoaffect oxy-char combustion, such as the interactions of theheterogeneous and homogeneous reactions, the multi-componentdiffusion within and around the porous char particle, as well asthe catalytic effects of coal ash minerals. Comprehensive modelsare needed for a better understanding.

2.7. Pollutants formation

Coal combustion in air results in the formation of a number ofregulated pollutants, such as NOx, SOx, mercury, and particulatematter including fly ash and soot. The formation of these pollutantsunder oxy-fired conditions has also been studied recently, and theessential findings in the literature are summarized with a focus onthe impact of replacing nitrogen with carbon dioxide.

2.7.1. Fly ash formationThe transformation of the mineral matters during coal

combustion depends on the char temperature and gas compositionwithin and around the coal particle. The mineral transformationand particle formation mechanisms during combustion are sche-matically shown in Fig. 44. Coarse fly ash particles are formedthrough coalescence and fragmentation of the non-volatile mineralcontent. Fine fly ash particles of submicrometer sizes are producedthrough two mechanisms [168]:

Fig. 44. Mineral transformation and particle forma

(a) Vaporization and condensation of heavy metals and organicallyassociated cations. The volatile species such as sodium (Na),lead (Pb), cadmium (Cd), and mercury (Hg) may undergo gas-phase chemical reactions and subsequently nucleate to formnew particles or condense on the surface of existing particles.

(b) Chemical transformation via reduction and oxidation reactions.Nonvolatile species, such as silica, with high melting temper-atures are released from the burning char particle in the formof suboxides via a reduction reaction at the inner surface of theparticle. These metals or mineral suboxides, which are morevolatile than the corresponding refractory oxides, diffuse awayfrom the parent coal particle, reoxidize, become supersaturatedand form particles by nucleation, followed by growth bycondensation and collision mechanisms.

Stam et al. [169] investigated the thermodynamic equilibriumcomposition of fly ash and slag in oxy-coal combustion and foundsimilar fly ash compositions in both air-fuel and oxy-fuelcombustion. Calculations indicated no significant change in thespeciation of trace element such as Hg, Se, Cr and As, even whenconsidering their higher concentrations due to flue gas recycle.Sheng et al. [170] experimentally studied the crystalline phase of flyash from a DTF using the X-ray diffractometer (XRD) approach.They observed similar main crystalline phases in the fly ash formedin O2/N2 and O2/CO2 combustion atmospheres, except that moreironmelts into the glass silicates and less is oxidized, because of thelower char temperature and the high CO concentration within thechar particles. The authors expected an increased slaggingpropensity of the fly ash under oxy-fuel combustion conditions.

Although the compositions of fly ash are similar in air-fired andoxy-fired conditions, two factors may affect the fine fly ash particlesize distributions [171] (also see Fig. 44). First, the lower particletemperature in oxy-fuel combustion with an air-like O2 level mayslow down the vaporization of volatile species, and thus inhibit theformation of submicrometer ash particles. The second factor is thehigher CO2 partial pressure, which may deter the chemical trans-formation rate of mineral oxides to mineral suboxides, and resultsin a slower particle formation rate. Suriyawong and coworkers[168] measured the number distribution of submicrometer fly ash

tion pathways during coal combustion [168].

Page 33: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214188

particles generated from PRB coal and observed a smaller totalmass of submicrometer ash particles, as well as a smaller geometricmean size in oxy-fuel combustion. The distribution approaches thatof air combustion with progressively increasing O2 mole fractions.Similarly, smaller PM2.5 and submicrometer particle yields werealso observed by Sheng et al. [171] in their oxy-coal combustionexperiment conducted in a DTF with oxygen concentration of 20%.

In short, the fly ash particle formation pathways are compli-cated. In addition to the dependence on bulk gas compositionwhenchanging from N2 to CO2 dilution, it is also influenced by the coaltype, the coal particle size distribution, the sulfur content, as well asthe operating conditions [172]. Particulate matter formation andtheir control strategies in oxy-fuel combustion require furtherinvestigations.

2.7.2. Ash deposit/slag formationFouling and slagging problems caused by coal ash impedes the

heat transfer in heat exchangers and may harm the operationsafety. Up to the date of this review, very few studies have beenpublished for ash deposit and slag formation propensity in oxy-coalcombustion. In this subsection, the fundamentals of ash depositionin coal combustion are reviewed, followed by a speculative analysisof possible trends under oxy-fired conditions.

The deposition rate of fly ash onto the combustor walls, in bothfouling and slagging, depends on the stickiness of the ash particlesas well as the walls. Richards [173] defined a sticky ash particle asa particle with a viscosity that is lower than the ash meltingviscosity. Thus, the melting temperature of the fly ash particlesdetermines whether the condition for stickiness is satisfied. Themelting temperature is in turn determined by the ash composition.However, though necessary, a high particle temperature is notsufficient because the molten minerals might not be exposed to theparticle surface. Li et al. [174] have found that there is a critical coalconversion of about 88% at which the molten ash is exposed to thesurface, which makes a particle sticky. This is consistent with thedrop tube experimental findings of Fryda et al. [175] who observeda higher deposition propensity of coal particles with a longerexposure to high temperatures in oxy-coal combustion with 30%oxygen concentration, when compared with air combustion.

On the other hand, a wall is considered sticky when a layer ofmolten slag exists [176] as shown in Fig. 45. This is only possible ifthe wall temperature is higher than the melting temperature of thedepositing ash which is dependent on the coal type and ashcomposition.

Since the dominant mechanism in deposition is inertialimpaction [177], based on the experimental data in Senda et al.

Fig. 45. Ash deposition onto a sticky (slag covered) or non-sticky wall.

[178]. Yong et al. [176] further proposed that the capture efficiencyis also dependent on the particle’s Weber number, besides either orboth of the particle and wall being sticky. This dimensionlessnumber compares the kinetic energy of the particles and theinterfacial surface tension energy between the particles and thewall and is given by:

We hParticle Kinetic EnergySurface Tension Energy

¼ rp v2pdpssp

(31)

where rp is the particle density, vp the normal component ofparticle velocity, dp the particle diameter and ssp the wall-particlesurface tension. The wall-particle surface tension is determinedby the surface tension of the sticky wall or of the sticky particle ifonly one of them is sticky. However, if both the particle and thewallare sticky, the sticky particle is always captured and if both theparticle and the wall are non-sticky, the particle is always reflected[176]. Therefore, in these cases, the wall-particle surface tension isimmaterial.

Based on the above considerations, the fouling or slaggingpropensity of an ash particle is dependent on the ash particlecomposition and conversion, as well as the particle’s Webernumber. The differences in ash deposition propensity between anoxy-coal and an air-coal combustion environment can thus besummarized as follows:

� As described in the previous section, a higher percentage ofiron species have been found in ash formed in oxy-coalcombustion than in air combustion [170]. According to theauthors, a higher iron concentration leads to a lower ashmelting temperature and an increased fouling/slaggingpropensity for the fly ash.

� Based on the discussions in Sections 2.1.3 and 2.5.1, with similaroxygen mole fractions, the combustion temperature is lower inan oxy-coal combustion environment and ignition is delayed.Therefore, carbon conversion of particles impacting the walltends to be lower. Thus, the capture efficiency is affected anddepending on the combustor design, ash particles may bedeposited further downstream from the burner inlet.

� Given the same momentum of the primary stream flow in theburner (which will be discussed in Section 3.3.3.1), the velocityof the particles that impact the combustor wall is reduced inthe oxy-coal environment. This results in a lower Webernumber for the particles and a higher capture efficiency, whichonce again agrees with the observations of Fryda et al. [175] inan oxy-coal combustion environment. However, this effectvaries with combustor design and is not expected to havea significant impact.

2.7.3. NOx formationComprehensive literature reviews on NOx emission and its

control in oxy-fuel combustion have been written by Normannet al. [179] and Toftegaard et al. [20]. Therefore, this paper onlyfocuses on the fundamentals of nitrogen chemistry and themechanisms and factors that impact in NOx formation in oxy-fuelcombustion.

2.7.3.1. Nitrogen chemistry. Fig. 46 illustrates the fundamentals ofnitrogen chemistry in combustion processes [179]. Since NO is themost important nitrogen-oxygen product of combustion, makingup most NOx emissions, and this paper confines the discussion toNO formation. NO is formed through three pathways: thermal andprompt NO from N2, and oxidation of fuel-bound nitrogen. Theformation of thermal NO is described by the extended Zeldovichmechanism:

Page 34: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Fig. 46. Mechanisms of NO formation and reduction [179].

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 189

N2 þ O4NOþ N (32)

Nþ O24NOþ O (33)

Nþ OH4NOþ H (34)

Reaction (32) is the rate-controlling step because of its highactivation energy, and temperatures above 1500 �C are required to

Table 10Research on NOx and SO2 emissions of air-coal and oxy-coal combustion.

Facility Scale NO recycle Emission in mg/MJ (convers

Thermal NOx

Air Oxy

IHI horizontal combustiontest facility [95,191]

1.2 MW Wet/Dry FGR 342 57

IHI vertical pilot scaletest facility [22]

1.2 MW Wet FGR w220e420 w75e120

Tokyo Institute ofTechnology Onedimensional premixedreactor [181]

Lab scale Doped NO e e

Nagoya UniversityEntrained flowreactor [180,184]

Lab scale O2/CO2 e e

Doped NO e e

initiate it. Prompt NO is initiated by the reaction between N2 andhydrocarbon radicals forming a volatile-N species (an intermediategaseous compound), which is either oxidized to become NO orreduced back to form N2 again. Prompt NO is typically of minorimportance and occurs predominantly under fuel-rich conditions.On the other hand, the conversion of fuel-bound nitrogencontributes to most of the NO formed during the combustion ofcoal. Fuel-N is split into volatile-N and char-N during devolatiliza-tion. The volatile-N transforms into either NO or N2, while char-Nreacts through a set of heterogeneous reactions along with theoxidation of char. NO can also be reduced to N2. Similar to theprompt formation, NO can be reduced by hydrocarbons to volatile-N under fuel-rich conditions (known as reburning). Furthermore, ifpresent in sufficient quantities, char could act as an importantcatalyst for the reduction of NO, either directly or by reaction withcarbon monoxide or hydrogen [179].

2.7.3.2. Lower NOx emission intensities in oxy-coal combus-tion. Emissions of NOx have been extensively investigated in labscale and pilot scale studies in both air-coal and oxy-coalcombustion environments, as summarized in Table 10. Almost allthe studies show lower NOx emission intensities (emission per unitthermal energy) in oxy-coal combustion than in air-coal combus-tion. For the sake of comparison, the observations in Table 10 arecompiled in Fig. 47. This figure indicates that all data points liebelow the dashed line which marks the case where emissionintensities are identical in oxy-coal combustion and air-coal

ion ratio) Experiment findings

SO2

Air Oxy

e e Lower NO emission in oxy-fuelcombustion compared toair-combustion.

w140-420 w100e300 NOx concentration increases 45%in oxy-coal combustion; Lower NOx

emission intensity (approximatelyone third of the total NOx in air-coalcombustion).SO2 concentration is three timesgreater in oxy-coal combustionthan in air-coal combustion;Sulfur emission intensity is twothirds of that in air-coal combustion.

e e Lower NOx formation in oxy-fuelcombustion due to three mechanisms:(i) reduction of recycled NO in thefurnace (dominant), (ii) interactionbetween fuel-N and recycled NO(secondary), and (iii) the effect ofincreased CO2 concentration(relatively insignificant).

e e NO formation is lower in oxy-fuelthan in air-fuel combustion; NOincreases with increasing O2 levelwith a peak at oxygen concentrationof 50%; NO emission increases withincreasing temperature in both air-and oxy-fuel environments; SO2

emission is almost independent ofenvironment, temperatureand O2 level.

e e Reduction efficiency of the recycledNO increases with increasingequivalence ratio, and recyclingratio (due to less reduced CHfragments in higher O2

environment) at fixed temperature

(continued on next page)

Page 35: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Table 10 (continued )

Facility Scale NO recycle Emission in mg/MJ (conversion ratio) Experiment findings

Thermal NOx SO2

Air Oxy Air Oxy

CANMET verticalcombustor [93]

0.21 MW O2/CO2,Wet/Dry FGR

w350 200e300, 100e200 (0.91) (0.75), (0.64) NO emissions are lower in oxy-coalcombustion; NO increases withincreasing O2 mole fraction; SO2

emission is independent ofcombustion environment.

[94] Partial FGR 211e269 68e233 e e SO2 concentration is 3e4 timeshigher than in air-coal combustiondue to recycle, but similar in termsof emission intensity; higher NOconcentration in furnace but lowerin terms of emission intensity.

University of LeedsDown-firedcombustor [182,183]

20 kW Na 340 278 (0.86e0.90) (0.86e0.90) Lower coal-N to NO conversion inoxy-fuel combustion; NO increaseswith increasing O2 mole fraction;SO2 emission is independent ofcombustion environment.

Doped NO e e e e Staging combustion reduces NO inboth air-coal and oxy-coalcombustion; without staging,68e82% of recycled NO is reducedin oxy-fuel combustion, more thanin air combustion (61e70%) for 7 coals.

RWTH Vertical pilot-scalefurnace [162]

100 kW Na 255 136e218 Lower NO in oxy-fuel combustion.

E.ON UK facility [189] 1 MW Dry FGR 97 25 431 277 Higher NO concentrations, while LowerNO emission intensities are observedunder oxy-fired conditions.SO2 concentration is 3.3 times higher foroxy-fired than for air-fired condition.SO2 emission intensity is 64% of thatunder air-fired condition.

IVD facility [189] 500 kW Wet FGR 191 67 531 394 Same trends of NO emission as above.SO2 concentration is 2.3 times higher foroxy-fired than for air-fired condition.SO2 emission intensity is 74% of thatunder air-fired condition.

Chalmers 100 kWtest unit [88]

100 kW Dry FGR 161 41e48 NO concentrations are comparable in airand oxy-fuel flue gas, but the emissionintensity is much lower in oxy-fuelcombustion.

a Without any flue gas recycle in the experiment.

Fig. 47. Comparison of the averaged NOx emission intensity in air-fired and oxy-firedcombustion of coal reported in the literature (see Table 10). B indicates test resultsfrom once-through NOx emission, , indicates test results with wet/dry flue gasrecycle. — shows the identical emissions, shows the trend without flue gas recycle,and shows the trend with flue gas recycle.

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214190

combustion. A reduction in NOx emissions is most pronounced inexperiments with FGR, with the slope of a least-square-fitted trendline being 0.22, which is substantially smaller than 1. Note that theCANMET results do not show as much NOx mitigation as the otherexperiments with FGR, and the lower NOx mitigation is mainlybecause their tests were performed under partial flue gas recycle(37%e45% of the flue gas bymass is recycled, and the remaining gaswas made up with supplied CO2), hence less NO was recycled andreburned. The reasons for the lower NOx emission intensities inoxy-coal combustion are discussed in greater detail in the followingsub-section.

2.7.3.3. Reasons for lower NOx emission intensities. The lower NOxemission intensities under oxy-fired conditions are due to thefollowing reasons:

(i) The lower N2 partial pressure in oxy-fuel combustion inhibitsthe formation of thermal and prompt NO.

(ii) CO2 concentrations in oxy-fuel combustion are substantiallyhigher than in conventional air combustion, which change theradical and gas compositions within and around the charparticles and thus influences the NO formation.

(iii) The NO in the recycled flue gas is reburned in the combustionzone.

Page 36: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 191

These reasons are discussed in greater detail as follows:To better examine the influence of the lower N2 partial pressure

on the formation of NO, Normann et al. [179] calculated thechemical equilibrium concentration of NO under both fuel-rich(l ¼ 0:9) and fuel-lean (l ¼ 1:1) conditions with increasing N2concentrations (between 0 and 10%). The results show that theequilibrium NO concentration is reduced dramatically at lower N2partial pressures, above 1500 �C in fuel-lean condition and above1700 �C in fuel-rich condition. This has implications for oxy-fuelcombustion, because the N2 impurity in the oxygen streamcontributes to NO formation. It also shows that fuel-rich condi-tions are favorable for NO reduction in oxy-fuel combustion, andthat staging technologies are still beneficial under oxy-coalconditions.

By conducting oxy-fuel combustion experiments without fluegas recycle, the impact of replacing N2 with CO2 could be exam-ined without the interfering effect of NO reburning. Studieswithout flue gas recycle show that, similar to NO formation in air-fuel combustion, the fuel-N conversion ratio in oxy-fuel combus-tion is a function of temperature, O2 concentration and equiva-lence ratio. For instance, Hu et al. [180] conducted NO emissiontests using a lab scale entrained flow reactor and observed that thefuel-N conversion rises with increasing temperatures and O2

concentrations, as well as decreasing equivalence ratio [180,181] inboth air-fired and oxy-fired coal combustion. Like that in the air-combustion, the conversion ratio decreases with combustionstaging [182].

Lower fuel-bound nitrogen conversions in oxy-fuel combus-tion were reported in all of these studies. The lower fuel-Nconversion ratio is suspected to be due to the different COconcentration. Higher CO is expected in and around the charparticle under oxy-fired conditions due to CO2 dissociation andthe gasification reactions, leading to a reduction in the amount ofNO formed during char combustion [148,149]. Okazaki et al. [181]investigated the influence of CO2 concentration on fuel-N NOformation using a flat methane-coal flame, and found that thefuel-N conversion ratio decreases slightly with increasing CO2concentrations. likewise, Liu et al. [183] found the same trend inan entrained flow reactor test, whereas the fuel-N conversion ina 30% O2/70% CO2 environment is reduced by a factor of 0.5e0.9below that in air-coal combustion. The discrepancy in the degreeof reduced conversion in these studies might be explained by thedifferent fuels used and the incomplete conversion of the coal inOkazaki et al.’s experiments.

Moreover, of all the NOx reduction mechanisms, research hasshown that reburning and reduction of the recycled NOwith FGR ina fuel-rich environment contributes the most. Okazaki et al. [181]studied the reduction of doped NO in an oxy-CH4 flame andfound a reduction ratio of 0.3e0.6 while varying the equivalenceratios. Comparing this with other mechanisms that lead to reducedNO formation in oxy-fuel combustion, the authors concluded thatthe reduction of recycled NO is the dominant mechanism.

Analogous to the NO reburning process in air-combustion withflue gas recycle, the reduction ratio is dependent on the coalreactivity, the equivalence ratio and the flue gas recycle ratio (or theO2/CO2 concentration). Higher reduction efficiencies have beenreported for fuel-rich conditions in the studies of both Okazaki et al.[181] and Hu et al. [184]. Similarly, as reported by Liu et al. [183] andKiga et al. [185], higher reduction efficiencies were achieved withstaged combustion where fuel-rich stoichiometry is maintained inthe volatile combustion zone near the burner. The differencebetween NO reburning in air-fired and oxy-fired environments,however, is consistent with the trend of fuel-N conversion: thereduction efficiency of recycled NO is higher in oxy-fired conditionsthan in air-fired conditions. Liu et al. [183] reported higher

reduction efficiencies of doped NO in oxy-fired conditions than inair-fired conditions for 7 bituminous coals, and they attributed it tothe same reason as the lower fuel-N conversion in oxy-fuelcombustion: higher CO concentrations within the pores of coalparticles and in the bulk gas in oxy-fuel combustion inhibit theformation of NO. However, the NO reburning efficiency has beenshown to be independent of temperature [184] and the concen-tration of recycled NO [181,183].

2.7.4. SO2 formationAs shown in Table 10, SO2 emissions in oxy-coal combustion

have been tested at the lab scale [180,181] and pilot scale[22,93,94,162,182,183]. These studies found that the fuel sulfurconversion ratio is in the range of 0.64e0.90, and the emissionintensity does not change significantly under oxy-fuel and air-fuelcombustion conditions. However, since partial NOx and SOx arerecycled with the recycled flue gas, NOx and SO2 concentrations inthe furnace are significantly higher than in air-fuel combustion.This may raise potential corrosion issues at the end of the flue gasduct if it is cooled down below its dew point [94]. The reader isreferred to references [20,186] for amore detailed review of the fateof sulfur under oxy-coal combustion.

2.7.5. CO concentration and emissionCO is produced from devolatilization, the oxidation of volatile

matter in the near-burner regions and char gasification reactions inthe char combustion zone. At high temperatures, the equilibriumCO concentration is governed by the following reversible reaction:

COþ 12O24CO2 (35)

Equilibrium of this exothermic reaction is shifted toward CO2dissociation at high temperatures. Hence, CO concentration is ex-pected to be higher in oxy-fuel combustion because of the higherCO2 partial pressure at the same temperature.

Zheng and Furimsky [187] calculated the COmole fraction in theequilibrium state in oxy-coal combustion and found that the COconcentration is 316 ppm at 1700 K when burning with 10% excessoxygen. This is substantially higher than that of air combustion,which has an equilibrium CO concentration of about 64 ppm at thesame stoichiometric ratio. Glarborg and Bentzen [188] experi-mentally investigated CO concentrations during methanecombustion in highly diluted environments with either N2 or CO2as the diluent, in a plug-flow reactor with well-defined tempera-ture profiles. Fig. 48 shows the CO concentration in the product gasat the exit of the reactor under lean, stoichiometric and richconditions in O2/N2 and O2/CO2 atmospheres. Substantially higherCO concentrations in the CO2 diluted environment than in the N2diluted environment were observed in all cases. Moreover, that theCO concentration increases from fuel-lean to fuel-rich conditions inboth diluent gases agrees with the trends of the equilibriumcalculations in [187]. An accurate prediction of the CO concentra-tion in the flame zone in oxy-fuel combustion becomes challengingand will be discussed in greater detail in Section 4.5.

Higher CO concentrations in the furnace have also been reportedunder oxy-coal conditions. Hjartstam et al. [88] observed that COconcentrations in the combustion zone near the burner, where thegas temperature is high and fuel-rich conditions prevail, are higherunder the OF27 (OF27 means 27% O2/73% CO2) and OF29 oxy-fuelconditions than under air-fired condition. Higher CO concentra-tions in the furnace were also reported in CANMET’s tests at similaror higher combustion temperatures [92,94]. Likewise, Rehfeldtet al. [189] found a significantly higher CO concentration in theoxygen lean flame region of oxy-Lausitz lignite combustion usinga 0.5 MW pilot scale test facility.

Page 37: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Fig. 48. Experimental data of CO concentration in the methane combustion productgas at the outlet of a flow reactor as function of temperature and stoichiometry, withN2 or CO2 as bulk gas. From the top, the graphs show the CO mole fraction under lean,stoichiometric, and fuel-rich conditions [188]. The symbol > indicates burning in N2diluent gas, while D and P indicate burning in CO2 diluent gas.

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214192

However, the high CO concentration in the combustion zonedoes not necessarily result in higher CO emissions because of itsnearly complete oxidation downstream at lower temperatures. Forinstance, different trends on CO emission in oxy-fuel and air-fuelmodes were reported in pilot scale tests: A lower CO emissionintensity (mg/MJ) in oxy-fuel mode was reported in the experi-ments conducted by Hjartstam et al. [88], whereas Rehfeldt [189]reported a generally higher CO concentration and emission inten-sity in oxy-fuel flue gas independent of the firing system, the coalrank and minor deviations in the overall stoichiometry under theoperating conditions in their experiments. In general, CO emissionmainly depends on the burn-out of the fuel, which is determined bythe burner performance and operating conditions such as stoichi-ometry. Managing CO emission in oxy-coal combustion would notbe a major issue if the combustion is well controlled and complete.

2.8. Summary

In this section, the physics and chemistry of oxy-coal combus-tion have been discussed in the order of the sub-processes that takeplace when a coal particle is burned in a CO2-rich environment. Inaddition to reviewing experimental findings, simplified modelswere applied to identify the impact of the distinct physical prop-erties and chemical effects of CO2 on the combustion characteris-tics. The causes and effects are summarized as follows:

� Heat transfer: Emissivity and absorptivity of oxy-fuel combus-tion flue gases are enhanced due to the higher partial pressure

of the triatomic gases, which results in different radiative heattransfer characteristics. Convective heat transfer coefficientsare also changed due to the different flue gas compositions.Flue gas recycle ratio is varied in the pilot scale oxy-combustionfacilities to obtain matched gas temperature profiles and/orheat fluxes. The optimized recycle ratio is dependent on thewet/dry recycle strategy, the fuel type and the heat exchangerarrangement. A matched distribution of the radiative andconvective heat transfer can be achieved at a slightly lowerflame temperature in oxy-coal combustion than in air-coalcombustion.

� Fuel Injection: The larger density of CO2 increases the drag forceduring the pneumatic conveying of coal particles if a dry recycleflue gas consisted mainly of CO2 is used as a carrier gas forpulverized coal. The smaller interfacial tension facilitates theatomization of the coal-water slurry, but only at high pressures.

� Particle heating and moisture evaporation: Calculations usinga heat transfer model show that the heating and moisture evap-oration rates in CO2-rich atmosphere in the mill and combustionzone are similar to those in air-coal combustion, given similarsurrounding gases and flame temperatures. Considering thepossibly lower combustion temperature in oxy-coal combustion,the heating time may be longer and ignition can be delayed.

� Coal devolatilization: Lab scale TGA and drop tube furnace testsunder different operating conditions indicate that the weightloss might be higher or lower in a CO2 environment than thoseunder the N2 conditions, depending on the gas temperatureand residence time. It is believed that the higher weight loss attemperatures above 1000 K in CO2 is due to the contribution ofgasification reactions, however, there is no evident influence ofthe oxy-fuel environment on the devolatilization process.

� Ignition delay: Lab scale experiments show that ignition delay islonger in O2/CO2 environments than in O2/N2 environments forseveral coal types. Results from a simplified simulation onmethane ignition delay with detailed reaction mechanismindicated that the high volumetric heat capacity of CO2 is themain cause of the homogenous ignition delay.

� Oxy-char combustion: The effects of CO2 on char consumptionrate are two-fold. At high temperatures and sufficiently high O2mole fractions where reactions are diffusion controlled, thechar consumption rate is lower due to the lower binary diffu-sivity of O2 in CO2; whereas at high temperatures and deficientO2 mole fractions, the char-CO2 gasification reaction canincrease the char consumption rate. The effect of pressure onthe oxidation and gasification reactions are discussed based onthe knowledge acquired in coal gasification studies.

� Pollutant formation: Experimental studies showed less sub-micron particulate matter in the fly ash distribution underoxy-coal combustion conditions when burning coal with anair-like oxygen mole fraction. Significantly lower NOx emissionper unit energy was reported in the literature, although NOxconcentrations in the flue gas may be comparable with[88,189,190], or 2 times higher than in the air-coal combustiondue to flue gas recycle [93,94]. The lower NOx emissionintensity is a result of the NOx reduction during flue gas recycle,the lower rate of thermal and prompt NO formation, as well asthe higher CO concentrations. Given similar combustiontemperatures, CO concentration is significantly higher in oxy-coal combustion than in air-coal combustion due to thethermal dissociation of CO2.

It should be noted that there are a few studies on pressurizedoxy-coal combustion. However, the fundamentals derived fromstudies on atmospheric oxy-coal combustion can help bridge theknowledge gap and aid in the development of pressurized systems.

Page 38: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Fig. 49. Maximum dilution ratio/minimum oxygen concentration in the oxidizerstream for methane-O2-dilution gas ignition in the WSR with a residence time of 0.1 s.

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 193

3. Stabilization of oxy-coal combustion

The challenges to maintaining oxy-coal combustion stabilityhave been reported in early pilot scale experimental studies[95,191]. These challenges stem from the lower adiabatic flametemperature, the delayed ignition and the lower burning rate ofcoal particles in a CO2 diluted environment, among other funda-mental issues discussed above.

Muñiz and Mungal [192] measured the standoff distance andvelocity of lifted methane and ethylene turbulent diffusion flamesand suggested two criteria for flame stabilization: (1) the compo-sition of the fuel/oxidizer mixture must be within the flammabilitylimits; (2) the local flow velocity must be near the premixedlaminar burning velocity (SL). In this section, the flammabilityconstraints as measured by the preheat temperature and dilutionratio under N2 and CO2 diluted combustion conditions are inves-tigated using a simplified WSR model. Possible region of stablecombustion in O2/CO2 combustion are calculated and comparedwith that under O2/N2 conditions. Regarding the second criterion ofstability, the characteristics of laminar burning velocity in oxy-fuelcombustion is also reviewed, as well as its implications on practicaloperations. Recent studies and measures taken to improve thestabilization of oxy-coal combustion are introduced.

Fig. 50. Schematic diagram showing the flammable domain expressed in terms of fuel/O2 equivalence ratio f, preheated temperature of reactants, and dilution ratio.

3.1. Flammability under oxy-fuel conditions

Prior to discussing oxy-fuel combustion stabilization, it is usefulto quantitatively sketch out the flammability domain in terms ofthe operating conditions. In conventional air-fuel combustion,flammability limits are frequently defined as the lower and upperbounds of fuel percentage by volume in the mixture, betweenwhich steady flame propagation occurs [109]. The use of almostpure oxygen and a fraction of recycled flue gases adds one extradegree of freedom to the problem, i.e., the dilution ratio. Therefore,three key factors are considered in the investigation of oxy-fuelflammable domains: the equivalence ratio, f, the reactant preheattemperature, T0, and the dilution ratio, DR. The DR is defined as theratio of molar flow rate of the dilution gas to the molar flow rate ofoxygen in the oxidizer gas stream:

DR ¼ _nDiluent gas= _nOxygen (36)

To gain some insight into flammability under oxy-combustionconditions, the flammable region of methane in O2/N2 and O2/CO2 is calculated using the well-stirred reactor (WSR) model andthe detailed reaction mechanism GRI-Mech 3.0 [139] with anadiabatic thermal boundary condition in CHEMKIN [138]. Theresidence time of the reacting methane-oxygen-dilution gas (N2 orrecycled dry CO2)mixture is fixed at 0.1 s. Blowout occurs when f isoutside the flammability limits at a given T0 and DR, or when DR istoo high given T0 and f.

Fig. 49 shows the critical dilution ratios at various preheattemperatures and equivalence ratios under O2/N2 and O2/CO2conditions. The curves shown are boundaries of the flammableregion indicating the maximum dilution ratio that corresponds toa specific preheat temperature and equivalence ratio. Flammabilityis assured for any operating conditions under the curve in the WSRwith a residence time of 0.1 s. The maximum dilution ratio isattained at an equivalence ratio close to stoichiometric methane/oxygen conditions (f ¼ 1) for all initial temperatures, and theflammable region expands when increasing the preheat tempera-ture. For the same T0, blow out occurs at lower dilution ratios foroxy-fuel combustion, resulting in a smaller flammable regionwhencompared with combustion under O2/N2 conditions. The smallerflammable region is caused by the lower combustion temperature

resulting from the higher molar heat capacity of CO2 (also see 2.1.3)and the lower reaction rate caused by the chemical effect of CO2(also see 3.2.1).

Based on the results of Fig. 49, the flammable regions with dilu-tion gases of N2 or CO2 are illustrated schematically in a DR� f� T0space as shown in Fig. 50. This flammable region is extended toa higher dilution ratio when the reactants are preheated and whenthe equivalence ratio is closer to stoichiometric. The envelope of theflammable region indicates the critical operating conditions beyondwhich blowout occurs. Combustion becomes less stable as theoperating conditions approach the flammable surface. The mostinteresting finding is that the flammable region shrinks significantlywhenchanging thedilutiongas fromN2 toCO2, exhibitinga tendencytoward destabilization in oxy-fuel combustion.

3.2. Laminar flame propagation

Pulverized coal volatiles are burned under conditions of partialpremixed or nonpremixed (when the primary stream is a recycledflue gas) turbulent flames in vicinity of the burner. The rule foranchoring a flame in both laminar and turbulent flow, at conditionsaway from blowout, is to match the local flame speed with the localunburned stream velocity [109]. Studies on conventional air-fuel

Page 39: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214194

combustion have shown that the laminar burning velocity is animportant factor in determining the blow off/blow out velocityunder premixed [193] and nonpremixed flame conditions[192,194,195]. Kalghatgi [195] suggested the following correlationto estimate blow out flow velocity, vblowout, for jet flames:

vblowoutSL;max

�rurN

�1:5

¼ 0:017ReH�1� 3:5$10�6ReH

�(37)

where SL;max is the maximum laminar burning velocity near stoi-chiometric ratios for hydrocarbons. The Reynolds number, ReH , isgiven by

ReH ¼ ruSL;maxHmu

(38)

The characteristic length, H, is the distance along the burner axiswhere the mean fuel concentration has fallen to its stoichiometricvalue and can be estimated by

H ¼ 4

YF;uYF:stoic

�rurN

�1=2�5:8

�dj (39)

where dj is the diameter of the jet. This correlation was alsoshown to be valid for methane and propane jet combustion withpartial CO2 dilution up to 60% in the fuel stream. As discussedabove, the laminar burning velocity, SL, characterizes the stability inpractical combustion facilities, and hence it is worth reviewing theimpact of high CO2 concentration on it.

The laminar premixed burning velocity of coal/O2/diluent gascan be measured experimentally. Fig. 51 shows a schematicdiagram for flame propagation in a coal particle suspensionproposed by Taniguchi et al. [137]. A smaller number of particlesare first ignited at t ¼ t0, and their volatile flame heats up thesurrounding coal particles at t ¼ t1. Next, these heated particlesare devolatilized, followed by the ignition of these particles whenthe volatiles reach a certain concentration after an elapsed time ofDt. These processes repeat themselves and the spread of this chainreaction is the so-called flame propagation in coal suspension. It isexpected that the flame propagation process is strongly influencedby heat transfer, coal devolatilization kinetics and reaction kineticsof the volatiles. To get a better understanding of the flame prop-agation velocity, these factors are examined in the followingsubsections.

3.2.1. Effect of physical properties and chemical kineticsWhen replacing N2 with CO2 as a diluent gas, a change in the

laminar burning velocity is expected. To study the factors governing

Fig. 51. Model of flame propagation in pulverized coa

burning velocity in an idealized environment, a simplified onedimensional scaling analysis proposed by Spalding [109,196] forgaseous fuel combustion is used, which couples the essentialphysics of heat transfer, mass transfer, chemical kinetics andthermodynamics:

SL ¼� 2aðnþ 1Þ _m

000

Fru

�1=2(40)

where a is the thermal diffusivity, n is the mass ratio of oxidizerto fuel, ru is the density of unburned gas, and _m

000

F is the fuelconsumption rate per unit volume. Eq. (40) captures the depen-dence of the laminar burning velocity (SL) on the thermodynamicand transport properties of the gas mixture, and can be used to gainqualitative insight into the process shown in Fig. 51. These factorsare summarized as follows, followed by a discussion in greaterdetail:

� Thermal diffusivity: The thermal diffusivity is the thermalconductivity, k, divided by the volumetric heat capacity, rcp.CO2 has a thermal diffusivity that is 0.64 times lower than N2(see Table 3). Therefore, the temperature of CO2 responds lessrapidly to a change in temperature boundary conditions [81],resulting in a lower burning velocity. The dependence onthermal diffusivity reflects the impact of heat transfer on theflame propagation process. The fact that a higher flame prop-agation velocity is expected in a diluent gas with higherthermal diffusivity has also been observed in gaseous fuelcombustion. Law [132] has shown that the laminar burningfluxes of the methane-O2-diluent flame follow the same orderas the diluent gas thermal diffusivity (He > Ar > N2) (seeTable 7).

Though the laminar burning velocity is dependent on thermaldiffusivity, Shaddix and Molina [34] suggested that stability iseventually independent of the thermal diffusivity. They performeda dimensionless analysis on the extinction theory, and showed thatthe extinction process is characterized by the inverse of the

chemical timescale sch ¼ a

S2L, and a is canceled because SL is

proportional to the square root of the thermal diffusivity (seeEq. (40)).

� Heat capacity: A higher heat capacity per unit volume, rcp,decreases the fuel consumption rate, _m

000

F , because of the lowerreaction temperature. The dependence on rcp reflects theinfluence of longer ignition delay and lower reaction rate on

l air mixture, proposed by Taniguchi et al. [137].

Page 40: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Fig. 53. Flame propagation velocity in different atmospheres and coal concentrations,the O2 concentration is 40% [197].

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 195

the flame propagation process. It should be noted that theeffect of higher heat capacity on SL is much larger than thermaldiffusivity, because it reduces the reaction temperature thathas exponential dependence on reaction rate _m

000

F through theArrhenius law. The effect of higher volumetric heat capacity onlonger ignition delay has been previously discussed in Section2.5.1. Likewise, other factors that influence the ignition delaydiscussed in 2.5.1 have similar impacts on the flame propaga-tion velocity, and will not be repeated here.

� Chemical kinetics: In addition to its distinct physical properties,CO2 is also active chemically and can change the radicalconcentration and hence the burning velocity.

The lower flame propagation velocity of pulverized coalcombustion in CO2 has been reported by Kiga et al. [185] and Sudaet al. [197] using the microgravity drop shaft facility in the JapanMicrogravity Center (JAMIC). The experiment on flame propagationvelocity within a coal-dust cloud was carried out in a microgravityenvironment to obtain a spatially homogenous distribution of coalparticles and to minimize the effect of natural convection andbuoyancy. In this experimental set-up, a spherical flame front wasobserved, from which the flame propagation velocities weremeasured. Fig. 52 shows snap shots of the flame propagationprocess in a coal-O2-diluent mixture: (a) for a higher volatile coal Ain an O2/N2 environment; and (b) and (c) for a lower volatile coal Cin O2/N2 and O2/CO2 environments, respectively. It can be seen thatthe flame propagation velocity depends on both coal devolatiliza-tion and the diluent used. The fact that the coal flame propagationvelocity increases with higher volatile contents is consistent withthe experiments reported in [137]. Furthermore, a comparisonbetween Fig. 52 and Fig. 53(c) shows that the flame propagationvelocity is significantly slower, and the brightness is also substan-tially lower, in CO2 with the same coal and oxygen concentrations.

The measured flame propagation velocities as a function of coalconcentration in the suspension are shown in Fig. 53 for threedifferent coals burning with the same oxygen concentration. Thepropagation velocity attains a maximum value at a moderate coalconcentration, where the maximum is found by balancing theopposing effects of the reaction heat release and the volume of coal

Fig. 52. Flame propagation in pulverized coal clouds suspension. Coal particle diam-eter: 53e63 mm, O2: 40%, dilution gas: 60% [197].

(additional heat capacity) that has to be heated. As observed in thesnap shots, both the higher volatile coal (A) and the lower volatilecoal (C) have much slower flame speeds in an oxy-fuel combustionenvironment.

To determine the effect of the diluent gas physical properties,the flame propagation velocities in three diluent gases (nitrogen,argon and carbon dioxide) were further investigated in [197] andthe results are shown in Fig. 54. The flame propagation velocitiesfollow the following order: SLðArÞ > SLðN2Þ > SLðCO2Þ. This isa result of the thermal diffusivities, a, of the diluent gases being inthe same order of aAr > aN2

> aCO2, and the heat capacities (per

mole) in the inverse order rcp;Ar < rcp;N2< rcp;CO2

(see Table 7).Based on the scaling analysis discussed above, both the increase inthermal diffusivity and the decrease in heat capacity result inincreasing the burning velocity.

The last factor, chemical effect of CO2, also plays a role but not asbig as the physical properties. Liu et al. [198] numerically investi-gated the burning velocities of methane and hydrogen under air-fired and oxy-fired conditions using a CHEMKIN-based model andGRI-Mech 3.0 mechanism. In their study, the kinetic effect of CO2was isolated by implementing a fictitious specious FCO2 that hadidentical thermal and transport properties as CO2 but was notchemical active. The results for the laminar burning velocities ofCH4 and H2 with diluent gases of N2, CO2 and FCO2 are shown inFig. 55. Burning velocities in three cases were calculated: fuel/O2/

Fig. 54. Impact of dilution gas and O2 concentration on coal flame propagation velocity[197].

Page 41: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

a b

Fig. 55. Burning velocities of methane and hydrogen mixtures at increasing equivalence ratios. Oxygen mole fraction in the oxidizer is kept at 21% in all cases. Also plotted are theexperimental data of Zhu et al. [247] in methane mixture and Westbrook [248] in hydrogen mixtures as filled symbols [198]. The symbol -B- indicates results using CO2, -D-indicates results using FCO2.

Fig. 56. The laminar burning velocity (SL) of methane combustions in O2/N2 and O2/CO2 conditions at different equivalence ratios (f) and elevated initial temperatures,experimental data are cited from [199,200]. Unburned gas are at 300 K, 600 K, 900 Kand 1200 K, 1 atm.

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214196

N2, fuel/O2/CO2 and fuel/O2/N2/CO2 in which 30% N2 in air wasreplaced by CO2. Since the FCO2 is “inert” chemically, the differencebetween fuel/air and fuel/O2/FCO2 shows the effect of the physicalproperties on the burning velocity, while the difference betweenfuel/O2/FCO2 and fuel/O2/CO2 indicates the chemical effect of CO2.Further analysis found that the dominant reaction pathway for thechemical role of CO2 is as follows:

COþ OH4CO2 þ H (41)

CO2 competes for the H radicals through the reverse reaction inEq. (41) with the most important chain branching reaction:

Hþ O24Oþ OH (42)

Thus, the presence of CO2 significantly reduces the concentra-tions of important radicals, i.e. O and H radicals, leading toa reduction of the fuel burning rate [198]. Fig. 55 indicates thatCO2

0s chemical effects play some role in reducing the burningvelocity when replacing N2 with CO2. However, its physical prop-erties have a more significant contribution.

3.2.2. Impact of operating conditionsSince the lower flame propagation velocity in oxy-coal

combustion may result in a lower blow out velocity, maintainingthe same operating conditions as in air-fuel combustion mayweaken the flame stability. In this section, the laminar burningvelocities in fuel-oxygen-diluent gas combustionwith nitrogen andcarbon dioxide as diluent gases are calculated to investigate theeffects of operating conditions. To simplify the computation,methane is once again used as a surrogate for the volatiles, and thedevolatilization process is neglected. The simulation is performedusing the PREMIX package of CHEMKIN [138] while employing theGRI-Mech 3.0 detailed reaction mechanism [139].

Fig. 56 shows the predicted SL in methane-O2-N2 and methane-O2eCO2 mixture with an oxygen concentration of 21 vol%. Experi-mental results at room temperature [199,200] are also shown for

comparison. Similar to the results obtained by Liu et al. [198] inFig. 55, the laminar burning velocities in oxy-fuel conditions aresignificantly lower than in air combustion: the maximum SL isabout 37 cm/s in the CH4-O2-N2 mixture near the stoichiometrycondition, as opposed to an SL of about 3e4 cm/s in theCH4eO2eCO2 mixture, both with initial temperatures ofT0 ¼ 300 K . The simulation also predicts the trends for a broaderrange of operating conditions, in particular with higher initialtemperatures. Preheating the reactants results in faster reactionkinetics and generally improves the flame propagation (higher SL).Oneway to achieve a higher preheat temperature is bymixing withhot exhaust gas. However, this also results in a higher dilution ofthe oxidizer stream.

Fig. 57 shows the impact of the dilution ratio on the calculated SLat f ¼ 1 (also shown as a function of oxygen concentration in the

Page 42: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Fig. 57. The laminar burning velocity (SL) of stoichiometric methane/O2 combustionswith N2 and CO2 used as dilution gases at increasing dilution ratios. Unburned gas is at25 �C, 300 �C and 600 �C, 1 atm, all calculations are at f ¼ 1.

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 197

oxidizer stream on the top axis) for methane-O2-N2 and methane-O2eCO2 mixtures. SL decreases with increasing dilution ratio forboth dilution gases, and as expected, the curves for the two dilutiongases meet at an oxygen mole fraction of 100%, where they bothreach the full oxy-fuel combustion regime. The results indicate that,given the same initial temperature, a higher oxygen concentrationis required in oxy-fuel combustion to match the laminar burningvelocity of air combustion; or a higher initial temperature isnecessary for a fixed dilution ratio.

Results in this section have practical implications on the stabi-lization of oxy-fuel combustion. Flame stability can be improved bymatching the primary stream velocity with local turbulent burningvelocity. This may be achieved by enriching oxygen locally,enhancing hot burned gas recirculation, and increasing thecombustion temperature at lower FGR, and these measures will bediscussed in greater detail in the following section.

3.3. Stabilizing oxy-coal combustion

As discussed in Section 3.1 and 3.2, narrower flammability limitsand lower laminar burning velocity may cause combustion desta-bilization in oxy-coal combustion. However, operating in the oxy-fuel regime provides more flexibility because the oxygen concen-trations in the oxidizing streams are not restricted to 21%. Thisopens up opportunities for advanced oxy-coal burner designincluding the following variables:

� Flue gas recycle ratio (overall oxygen mole fraction): A lowerrecycle ratio or dilution ratio favors a higher flame tempera-ture, broader flammability limits, and laminar burning velocityas discussed in Section 3.1 and 3.2, but its lower bound islimited by material constraints in the heat exchangers.

� Oxygen concentrations in the primary, secondary and tertiarystreams: Coal burners employ a primary stream inwhich coal isdeliveredwith a carrier gas, a secondary, and at times a tertiary,stream in which oxidizer is delivered. The oxygen concentra-tion in different streams can be independent of one another.Dry or wet flue gas is usually used as the primary (carrier) gasin oxy-fuel combustion, but the choice should be based on anoptimization for flame stabilization. Different oxygen concen-trations can be used in different streams in a staged combus-tion scenario for stabilization and NOx control, sometimespartial pure oxygen can also be injected.

� The distribution of gas volumes in the different streams and theirmomenta: Since the overall volumetric flow rate is lower inoxy-fuel combustion, the dimensions of the burner should becarefully designed to achieve an optimal flowfield for mixing.

� The preheat temperature of different streams: As shown in theprevious section, preheating the reactant streams can improveflammability. Preheating the primary and secondary gasstreams in conventional air-fired PC boilers using an air-preheater is widely applied for energy saving purposes. Thesame strategy can be utilized in oxy-coal combustion toimprove combustion stability.

In recent years, studies have been conducted to investigateburner retrofit and new burner design in oxy-fuel combustion inorder to achieve better mixing of the fuel, oxidizer and hot fluegases. Some of these efforts are summarized in Table 11, witha selection of typical burner configurations shown in Fig. 58. Flamestabilization and burner modifications will be discussed in detail inthis section.

3.3.1. Optimized flue gas recycle ratiosFrom the discussion in previous sections, it can be concluded

that combustion characteristics, such as ignition, flame tempera-ture, burning velocity and char consumption, are improved athigher O2 fraction. Therefore, reducing the flue gas recycle ratio andthereby increasing the oxygen concentrations in the primary,secondary and tertiary streams will lead to stabilization of oxy-coalcombustion.

Liu et al. [182] reported an unstable oxy-coal combustion ina 20 kW down-fired furnace when operating the burner streams at21% O2/79% CO2. A dark volatile flamewas observed downstream ofthe burner, while coal particle ignition and peak temperature wereboth delayed. According to that study, the O2 mole fraction in themixture of O2/CO2 has to be raised to 30% or higher tomatch the gastemperature profile of oxy-coal combustion to that of air-coalcombustion. Under this 30% O2/70% CO2 condition, the measuredresidual carbon in the ash was also reduced, which indicates higherburnout efficiency than in air-fuel combustion. Likewise, in the oxy-coal combustion experiments performed using a German lignitecoal in the Chalmers 100 kW test unit by Hjartstam et al. andAndersson et al. [127,128] (see Fig. 58(c) for the burner structure),improved combustion stabilities were achieved when maintaininga stoichiometric fuel-oxygen ratio and increasing the oxygenconcentration in the staging streams from 25% to 29% (a recycleddry flue gas with 30 vol% oxygen was used as carrier gas for allcases). It was found that stable combustion could not be establishedwith O2 mole fractions lower than 25% in the feed gas streams. Incontrast, the cases with oxygen fractions of 27% and 29% have themost stable flames and are least sensitive to reduced stoichiometricratios.

3.3.2. Oxygen feeding strategiesThere is more flexibility in supplying oxygen to the combustion

zone in oxy-coal combustion than in air-coal combustion. Given anoverall fixed stoichiometric ratio (SR) and flue gas recycle ratio(FGR), the oxygen feed can be split and mixed with differentstreams at arbitrary concentrations, or injected independentlythrough oxygen nozzles. However, not all permutations of oxygenfeeding strategies lead to a stable attached flamewith lowpollutantemission, or to safe burner operation. Different partitioning strat-egies for feeding oxygen are discussed in this section.

3.3.2.1. Pure oxygen injection. Pure O2 injection has been intro-duced in several new designs and retrofits of oxy-coal burners tostabilize combustion at a favorable location in the furnace and to

Page 43: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Table 11Oxy-coal burner design: experiments and CFD modeling in the literatures.

Institution (Burner name) Retrofit/New Burner configuration (Diameter in mm) O2 nozzle Primary, secondary and tertiary stream Combustion performance

IHI (IHI horizontal combustiontest facility) [95,185,191]

Retrofit 1.2 MW, swirl burner, primary outlet D77,secondary outlet D135, w/O2 nozzle D50

Up to 15% of theoverall fed O2

1st: Pf coal, RFG2nd: Swirling (S ¼ 0.59e0.65), RFG/O2,O2 content 25%e30%

Unstable flame and higher UBC w/opure O2 injection

CCSD and IHI (IHI vertical pilotscale test facility) [207]

Retrofit 1.2 MW, swirl burner, simplification onburner structure w/o quarl or oil gun

N/A 1st: Pf coal, RFG2nd: Swirling (S ¼ 0.2), RFG/O2,O2 content 23%e27%

Ignition delay due to higher momentumratio of primary and secondary stream,as well as less swirl flow effect

CANMET(Burner A, B and C) [201,249]

New design 0.3 MW, swirl burner, primary, secondary,and tertiary stream, w/pure O2 nozzle

Up to 57% of theoverall O2

1st: Pf coal, RFG2nd: Swirling (S ¼ 1-2), RFG/O2,O2 content 28%3rd: unknown

Swirl number and configuration effectswere studied using a CFD modeling

Graz University (over-stoichiometric and under-stoichiometric burner) [202,203]

New design 3 MW, Swirl burner, w/quarl In the form oftertiary stream

1st: Pf coal, RFG2nd: Tangential entry swirling, RFG/O2,O2 content unknown3rd: Tangential entry swirling, pure O2

in over-stoichiometric burner

Combined over- and under-stoichiometricburners to reduce FGR ratio to 50%

Hitachi modeling (Low NOx

DS burner) [208,209]Retrofit 1 MW, swirl burner N/A 1st: Axial vane swirling, retain volume

flow rate as air combustion, RFG/O2

content unknown2nd: Axial vane swirling, retainmomentum as air combustion, Swirling,RFG/O2, O2 content w30%3rd: same as secondary stream

With 75.1% wet FGR ratio, a similarflame shape (Temp and velocity) wasobserved using CFD modeling

Chalmers University ofTechnology [85e88]

New design 100 kW, swirl burner, fuel lance,primary/secondary streamw/pure O2 lance,

Concentric O2 lance,but not used for thepublished studies.

Fuel lance: Pf coal, RFG/O2, O2 content 30%1st: Swirling (S ¼ 0.79), Pf coal, 30% dryRFG and O2, O2 content 25, 27 and 29%2nd: Swirling (S ¼ 0.21), 70% dry RFGand O2, O2 content 25, 27 and 29%

Flame cannot be established at OF21,while it is stable at OF27 and OF29conditions; Similar temperaturedistributions in OF25 and air combustion;OF29 may cause ash melting problem

University of Leeds [182,183] Retrofit 20 kW, swirl burner, quarl D140 N/A 1st: Non-swirling, Pf coal, air or O2/CO2,

O2 content 21%e30%2nd: Swirling, O2/CO2, O2 contents 21e35%3rd: Two opposing tubes as tertiaryoxidant injectors

Unstable flame with OF21; O2 has tobe increased to 30% or higher to matchthe air-fired gas temperature profile

Siemens-REI [206] New design 0.3 MW, swirl burner, w/quarl, N/A 1st: Pf coal, air or RFG/O2, O2

content flexibleInner and outer 2nd: RFG/O2, O2

content flexible

Tested different retrofit strategies forprimary stream; stable combustionachieved w/low O2 content in primary

RWTH Aachen University(Burner A, Oxy-1 andOxy-2) [162,210]

New design 40 kW, swirl burner, w/quarl, threedifferent configurations

N/A 1st: Pf coal, CO2/O2, O2 content 19%2nd: Swirling (S ¼ 1.2), CO2/O2, O2

content 18e21%3rd: CO2/O2, O2 content 18e21%

Stable swirl flames at low O2 contentwere achieved in Burner Oxy-1 andOxy-2 because of well designed flow field

L.Chenet

al./Progress

inEnergy

andCom

bustionScience

38(2012)

156e214

198

Page 44: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Fig. 58. Schematics of various burner design/retrofit concepts for flame stabilization in oxy-coal combustion. (a) IHI 1.2 MW swirl burner with central oxygen nozzle [191]. (b)CANMET swirl burner with oxygen nozzle inside or between the primary and secondary streams [249]. (c) Chalmers swirl burner [86,88]. (d) Hitachi low NOx DS burner [208,209].(e) RWTH Oxy-2 burner [162,210].

Fig. 59. The gas temperature peak rises and moves toward the burner when injectingpure oxygen through the oxygen nozzle in IHI 1.2 MW swirl burner under oxy-coalcombustion conditions. The excess O2 injection improved the ignition and stabiliza-tion [95].

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 199

produce compact flames. Kimura et al. [191] conducted oxy-coalcombustion experiments with a 1.2 MWth horizontal IHI swirlburner. In the experimental set-up, recycled flue gas was used inthe primary stream for coal transport, while a mixture of recycledflue gas and oxygen was used as a secondary stream. The authorsreported an unstable, dark combustion zone with high unburnedcarbon (UBC) when the O2 concentration in the secondary streamwas less than 30%. However, when a central oxygen nozzle, asillustrated in Fig. 58(a), was used to directly inject O2 in thecenterline axial direction of the burner, the flame was stabilized.Moreover, the UBC in oxy-coal combustion was reduced to the air-coal combustion level when up to 15% of the overall O2 was injectedthrough the oxygen nozzle andmaintaining the O2 concentration inthe secondary stream at 30%. Likewise, further studies on the samefacility by Nozaki et al. [95] showed that direct O2 injectionimproved combustion stability and coal ignition when comparedwith that under operating condition with the same O2 content inthe combustion gas but without O2 injection (see Fig. 59).

In another burner designed by CANMET [201], as illustrated inFig. 58(b), 57% of the overall O2 was injected through the O2 streamarranged inside or between the primary and secondary streams toensure combustion stability. Likewise, a pure O2 stream was alsoused as a tertiary stream aside from the primary and O2/RFGsecondary stream to enable a stable ignition in the over-stoichiometric burner developed by Graz University of Tech-nology [202,203].

Recently, Praxair [204] reviewed their development of variousoxygen nozzle prototypes for flame stabilization and NOx control inoxygen-enhanced combustion, and proposed the applications ofthese nozzles to oxy-fuel combustion for CCS as well. The idea is todeliver oxygen to the volatile combustion zone, and in doing so, theflame stability and char burnout will be improved. Injectinga suitable amount of oxygen at the fuel-rich region in a staged lowNOx burner also reduces NOx emissions, whereas oxygen injectionabove the designed amount may result in oxygen “punching

through” the fuel-rich zone and thus increasing NOx emissions.Fig. 60 shows schematics of different oxygen nozzle designs withaxial, radial, combination and reverse injections. Some of theinjection strategies produce mild mixing, such as the axial, radialand combination injection, while other designs (reverse injection)aremore aggressive inmixing the oxidizer and fuel near the burner.

The performances of different injection directions wereassessed based on the gas temperature profiles and NOx emissions.Fig. 61 shows the flame temperature profiles when various oxygennozzles are used in the 1.2 MWth industrial scale burner facility ofAlstom Power. The results show the performance of oxygen nozzleswith 5% and 10% oxygen “replacement”, which is defined as thefraction of oxygen delivered directly to the furnace through thenozzle over the stoichiometric oxygen supply. The peak

Page 45: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Fig. 60. Schematic of nozzles used for direct oxygen injection [204].

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214200

temperature moves closer to the burner when using an aggressivemixing strategy (reverse nozzle). However, this strategy should beused with caution because excessive overheating of the burnerfront may occur if too much oxygen is delivered. On the other hand,a mild mixing strategy (combination, radial, and 45� angle towardflame) allows a broader range of oxygen injection while stillenhancing flame stability, reducing NOx, and improving the carbonburnout. From a comparison of the two oxygen replacementsstrategies, the flame temperature is observed to be significantlyhigher when the replacement rate is raised.

a b

Fig. 61. Effect of oxygen nozzle designs (combination, radial, reverse, and 45� angletoward flame) on flame temperature profile. (a) With 5% oxygen replacement, and (b)with 10% oxygen replacement [204]. Air indicates the baseline flame temperaturewithout oxygen nozzle.

3.3.2.2. Oxygen partitioning in different streams. In typical PCburners, pulverized coal is delivered to the combustion zone by theprimary stream, while the secondary and tertiary streams providethe remaining oxidizer needed for further combustion reactions. Inconventional air-coal combustion, the oxygen content in theprimary air provides oxidant for volatiles burning and stabilizes theflame downstream of the burner flow. The secondary and tertiarystreams are usually injected co-axially with the primary stream, atincreasing radii andwith swirl, to addmore oxygen to burn the chardownstream of the primary combustion zone. When retrofitted tooperate under oxy-coal combustion, mixtures of recycled flue gasand oxygen with various concentrations are used in the primaryand secondary streams. From a safety point of view, the use of dryrecycled flue gas in the primary stream can avoid auto-ignitionwhen transporting the coal to the burner. However, a low oxygenmole fraction in the primary streammay cause destabilization [88].Adding an appropriate amount of O2 in the primary stream witha concentration below that of air is beneficial to flame stabilitywithout compromising safety.

Some recent experimental studies have investigated the effect ofoxygen partitioning and shifting in primary and secondary streamson flame stabilization. Zhang et al. [205] studied the standoffdistances of co-axial turbulent diffusion oxy-coal flames witha 40 kW non-swirl burner at the University of Utah. When pure CO2is used as the carrier gas in the primary stream (oxygen partialpressure is zero), an attached flame cannot be achieved at anoxygen/fuel ratio near stoichiometry. Increasing the oxygen-feedrate (i.e. the oxygen concentration) in the secondary streamabove a stoichiometric ratio decreased the standoff distance andenhanced flame stability. As shown in Fig. 62(a), the transitionpoint for an attached stable flame is 52% oxygen concentration inthe secondary stream (and 43% in the overall feeding gas). It isworth noting that the stoichiometric ratio has to be much largerthan unity to obtain attached stable flames when using pure CO2 asthe primary stream. However, when shifting the oxygen supplyfrom secondary stream to the primary stream while maintainingthe overall oxygen concentration at 40% (i.e. a constant l ¼ 1.15), itis possible to stabilize the flame near the burner. Fig. 62(b) showsthat for oxy-coal combustion, detached flames were observedat primary oxygen concentrations up to 14.4%, while the flamebegins to attach to the burner at an oxygen concentration of 20.7%.The primary stream oxygen concentration required for flame

a b

Fig. 62. (a) The effect of secondary oxygen mole fraction on the stand off distance(with pure CO2 primary stream). (b) The effect of primary oxygen mole fraction on thestand off distance (with fixed stoichiometric ratio) under O2/CO2 and O2/N2 conditions,two preheat temperatures were tested (489 K and 544 K) in the oxy-coal mode. Citedfrom [205].

Page 46: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 201

attachment, however, is much higher than that under O2/N2conditions with the matched stream momenta, indicating thechallenge of flame stabilization in oxy-coal combustion. Moreover,preheating the secondary stream from 216 �C to 271 �C makes theflame more stable.

The attached flame is more likely to be achieved with swirlburners, in which stable combustion is possible while introducingno oxygen into the primary stream. Fry et al. [206] performed oxy-coal combustion experiments with a 0.3 MW Siemens-REI swirlburner using western bituminous coal from the Skyline mine. Theyfound that when the overall oxygen concentration is 27%, anattached flame can be maintained while decreasing the primarystream oxygen concentration to as low as 2.7%, which is the typicalresidual oxygen concentration in the dry recycled flue gas (mean-while the oxygen concentration in the secondary stream had to beincreased to 34.5% in order to match the overall oxygen/RFG ratio).The discrepancy in the experiments of Zhang et al. and Fry et al. arecaused by the different burner characteristics. More hot gas can berecirculated internally using swirl burners, and hence it is possibleto achieve better flame stability with low oxygen concentrations inthe primary stream.

3.3.3. Fluid dynamics considerationsThe stability of oxy-coal combustion is dependent to a large

extent on the flowfield in vicinity of the burners. Since the overallgas volume flow rate is reduced in oxy-coal combustion, themomenta of different streams, the swirl effects and the corre-sponding mixing of fuel, oxidizer and recirculated hot gas may bechanged accordingly. Therefore, the geometry of the oxy-burnersshould be carefully redesigned in order to ensure stablecombustion.

3.3.3.1. Matched mass flow rate, momentum or velocity. Khare et al.[207] numerically studied the combustion aerodynamics of an IHI1.2 MW burner. Given the same burner geometry, the authorspointed out that if the same gas velocity of the primary stream is

Fig. 63. Axial velocity distribution under air-fired and oxy-fired operating conditions in th(internal and external recirculation zone) [189].

maintained in order to keep coal particles in suspension, the massflow rate of the recycled flue gas fed to the secondary stream has tobe reduced because the total gas volume in oxy-coal combustion isreduced to 70%e80% of that in air combustion. Therefore, themomentum ratio of the primary to the secondary burner streamsmay change. For instance, they found that the primary to secondarystream momentum ratio was increased from 0.13 for the air-coalcase to 0.73 for the oxy-coal case when maintaining the sameprimary velocity. This in turn decreases the swirl effect and changesthe flow regime. Consequently, a greater ignition delay is predictedwith a longer flame stand-off distance in the oxy-fuel case. Asimilar phenomenon was reported by Rehfeldt et al. [189] wheninvestigating the flow patterns of air-fired and oxy-fired combus-tion modes in the IVD (Institute of Process Engineering and PowerPlant Technology, University of Stuttgart) combustion test facility.The reduced volumetric flow rate results in a significantly lowersecondary/primary stream momentum ratio and an altered flowpattern. Fig. 63 shows that a type-2 flame (with internal recircu-lation) in air-coal combustion changed to a type-0 flame (withoutinternal recirculation) in oxy-coal combustion due to the weakerswirl effect of the secondary stream. The measured O2 and CO2

concentrations also indicate weak internal recirculation anda different flame shape in oxy-coal combustion.

On the other hand, it is possible to stabilize combustion bymaintaining the burner flow momenta when retrofitting. Tiggeset al. [208,209] simulated the combustion field of the Hitachi LowNOx DS burner shown in Fig. 58(d) under air-fired and oxy-firedconditions. The O2 mole fractions in the secondary and tertiarystreams were increased to match the flame temperature with thatof air-coal combustion. When keeping a similar volumetric flowrate for the primary stream and maintaining the momenta for allother streams, similar flowfields and temperature distributionswere obtained.

In the investigation of oxy-fuel burner retrofit strategies per-formed by Fry et al. [206], themain concernwas the influence of theprimary stream flow characteristics on flame stability. For this

e IVD combustion test facility. The white areas indicate negative axial velocity values

Page 47: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214202

purpose, they fixed the stoichiometric ratio, i.e. the overall O2concentration, and the ratio of the secondary to tertiary streamflow rates, and tested three cases of primary stream flow configu-rations: matching themass flow rate (m), matching themomentum(p), or matching the velocity (v) with that of the primary air in air-fuel combustion. The primary stream velocities under these caseswere in the following order: vm < vp < vv. They observed that bymatching the primary gas-to-fuel mass flow ratio and primarystream momentum to that of air-coal conditions, a stable andattached flame that closely resembled the air-fired flame wasattained. Conversely, when oxy-coal combustionwas operatedwitha matched primary stream velocity, the flame was detached fromthe burner quarl. Recalling the discussion in Section 2.2.1, it isunnecessary to maintain the primary stream velocity for coalparticle transport. Therefore, the primary stream velocity can beoptimized while improving the flame stabilization.

The findings in these studies suggest that flame stability isdependent on the flow characteristics, which in turn are deter-mined by the design and operation of the burner. The primarystream velocity should be optimized for delivering the pulverizedcoal, while considering its influence on the mass balance and cor-responding momentum ratios of different burner streams in oxy-coal combustion. The combined effect of the burner streams onflame stabilization should be taken into account when designingthe burner. More research is needed to better determine the designprinciples.

3.3.3.2. Advanced burner design. Combustion stability can also beattained manipulating the combustor flowfield through advancedburner design. The use of swirling flow, bluff body, and quarlgeometry can create a central recirculation zone, while the inletstreams are heated by recycled hot flue gas [106]. Most studies sofar have focused on retrofitting air-fired burners and optimizing theoperating conditions. One of the few research studies on new oxy-burner designs, carried out by Toporov and coworkers [162,210],improved the design of the Oxy-1 and the Oxy-2 burners (also seeFig. 58(e)) by using in-flame measurements and CFD modeling. Asshown in Fig. 64, the quarl geometry was changed in Burner Oxy-1and further improvement was made in Burner Oxy-2 whichfeatures a bluff body. These geometry improvements increased theinternal hot product recirculation, forced the ignition point closerto the quarl and made a stable oxy-coal flame possible at O2concentrations as low as 18%.

3.3.4. Gas stream preheatingAnother method to stabilize oxy-fuel combustion is through

preheating the burner streams. This method provides anotherdegree of freedom when operating oxy-coal burners. While pre-heating an oxygen-carrying primary stream may raise safetyconcerns, other streams could be preheated by means of heatrecycle through gasegas heat exchangers or flue gas recycle athigher temperatures. As discussed previously, preheating thesecondary stream improved the flame stability in Zhang’s experi-ment [205] (see Section 3.3.2.2). Preheating the inlet gas streams is

Fig. 64. Schematics of burner A, Burne

not new to air-fired PC combustion. Combustion of pulverized coalwith highly preheated air, exceeding 1000 K, has been demon-strated [37,211]. This experience could be utilized when applyingthe gas stream preheating technology to oxy-coal combustion[40,42,212].

4. CFD modeling of oxy-coal combustion

CFD techniques have become the third dimension in fluiddynamics and combustion studies alongside analytical modelingand experimental diagnostics [213]. CFD provides a relativelyinexpensive (when submodels are used in connection withReynolds-averaged NaviereStokes (RANS) or when using coarsegrain large-eddy simulation (LES) models) and indispensable toolto perform comprehensive studies on the fluid flow, heat transferand chemical reactions in combustion. Currently, CFD modeling ofoxy-coal combustion utilize approaches and sub-models that aresimilar to those developed under air-fired conditions. With theaccumulated knowledge on the fundamental differences betweenair-fuel and oxy-fuel combustion, some effort has gone intodeveloping and validating sub-models for the new combustionenvironment.

A selection of the CFD simulation studies on oxy-fuel combus-tion is summarized in Table 12, which includes the sub-modelsused for turbulence, radiation heat transfer, char combustion andhomogenous reactions. Since the existing sub-models weredeveloped for conventional air-coal combustion, their assumptionsand approximations may not be valid in the CO2-rich environment.In the following sections, the development of CFD sub-model for anaccurate prediction in oxy-coal combustion is reviewed, and thefindings of these recent numerical studies are summarized.

4.1. Governing equations and physical properties

Coal combustion is typically modeled as a dilute two-phase(solid-gas) reacting flow using an Eulerian-Lagrangian approach.The mass, momentum and energy interactions between the gasphase and the solid particles are calculated using the “particle-source-in-cell” method [214] while updating the particle statealong a set of particle trajectories. For the gas phase, the equationsgoverning mass, momentum, species and energy are written inconservative form:

vðrFÞvt

þ v�rujF

�vxj

¼ v

vxj

GF

vF

vxj

!þ SF (43)

with F, t, u, GF and SF denoting Favre-averaged variables, time,velocity, diffusion coefficient, and source term, respectively [215].These equations are the same in both air-coal combustion and oxy-coal combustion. However, the thermodynamics and the transportproperties of the gas mixture are different.

In turbulent flows, the effects of turbulence on the transportproperties, such as the turbulent viscosity, turbulent mass diffu-sivity, and turbulent thermal conductivity, generally dominate the

r Oxy-1 and Burner Oxy-2 [210].

Page 48: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

Table 12Summary of CFD simulations and their sub-models for oxy-fuel combustion.

Author Simulated object Modeling approaches

Facility Fuel Code Turbulence Radiation Charcombustion

HomogeneousReactionmechanism

Chemistry-Turbulence

Wang et al. [32] BCL Subscale combustor Wage coal 1-DICOG (1-D) N/A Zone Method [252],transparent gas

C þ O2

C þ CO2

C þ H2O

Volatilescombustion

ChemicalEquilibrium

Khare et al. [207] IHI 1.2 MWth verticalpilot scale test facility

Coal A Fluent k� ε P-1WSGG

C þ O2 Volatilescombustion

ChemicalEquilibrium

Nozaki et al. [95] IHI 1.2 MWth horizontalcombustion test facility

Coal A/B VEGA-3 k� ε Multi-flux Radiationmodel [253]Three-gray-gas model

C þ O2

C þ CO2

C þ H2O [254]

Volatilescombustion

EBU

Chui et al. [201] CANMET 0.3 MWth VCRF Western Canadiansub-bituminous coal

CFX-TASCflow Standard k� ε N/A C þ O2 Volatilescombustion

EBU

Rehfeldt et al. [189] E.ON 1 MWth horizontalfiring facility and IVD500 kWth down firing facility

Tselentis coal andLausitz lignite coal

Fluent Standard k� ε DO C þ O2

C þ CO2

N/A N/A

Toporov et al. [162] RWTH Aachen U test facility Rhenish lignite Fluent k� ε DOWSGG

C þ O2

C þ CO2

C þ H2O

VolatilebreakupCO and H2

burning [255]

Finite Rate/EddyDissipation

Chen et al. [80] ISOTHERM PWR� 5 MWthpressurized test facility

Bituminous coal Fluent Realizable k� ε, k� u DO WSGG C þ O2

C þ CO2

C þ H2O

Modified JL Finite Rate/EddyDissipation

Andersen et al. [233] 100 kW down-fired furnace [85] Propane Fluent Realizable k� ε P-1 N/A WD [229] andModified WD [233],JL [230] andModified JL [233]

EDC

Vascellari et al. [227] IFRF 2.4 MW furnace Gottelborn hvBp coal Fluent Standard k� ε P-1 C þ O2

C þ CO2

C þ H2O

Volatile decomposition,tar partially oxidation,Modified JL [239]

EDC

Muller et al. [215] IFK 0.5 MWth test facility Lausitz lignite AIOLOS Standard k� ε DOLeckner’s model [256]

C þ O2

C þ CO2

C þ H2O

JL [230] EDC

Nikolopouloset al. [224]

330 MWe PC boiler inMeliti power plant, Greece

Lignite fromAchlada mine

Fluent Standard k� ε DOEWBM

C þ O2

C þ CO2

Volatile combustionand CO burning

Finite Rate/EddyDissipation

Edge et al. [217] 0.5 MWth Air- and oxy-firedcombustion test facility withDoosan Babcock triple-stagedlow NOx burner and IFRFAerodynamicallyair-staged burner

Coal A and B Fluent RNG k� ε and LES DO WSGG/FSK NA Volatile combustionand CO burning

EDM

L.Chenet

al./Progress

inEnergy

andCom

bustionScience

38(2012)

156e214

203

Page 49: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214204

laminar transport properties, resulting in a relatively insignificantcontribution from the laminar properties. On the other hand,thermodynamic properties such as density and heat capacity arealways important, regardless of whether it is a laminar or a turbu-lent flow.

Chen et al. [80] have shown that the gas phase specific heatdistributions in oxy-fuel and air-fuel conditions are substantiallydifferent using a CFD simulation of a pilot-scale combustor atelevated pressure. In their study, a slightly higher O2 mole fractionwas used in the oxy-fired oxidizer stream (22.2%). The same swirlnumber was used for both cases, guaranteeing a similar overallflowfield (shown in the Fig. 65(b)) under oxy-fuel and air-fuelconditions. Fig. 65(a) shows that the specific heat distributionfor the oxy-fuel environment is in the range of 1300e1800 J/kg-K,significantly higher than for the air-fuel environment(1100e1400 J/kg-K), because of the higher concentration of CO2and H2O in the former. The higher specific heat under oxy-fuelconditions results in significantly lower combustion tempera-tures (shown in the Fig. 65(b)). Devolatilization and char burningare delayed in the oxy-fired case as a result of the lowertemperature.

Fig. 65. The predicted specific heat (cp) field and resulting temperature distribution in a prespecific heat (J/kgK) distribution; (b): streamlines are superimposed over temperature field

4.2. Turbulent flow

Some of the most important phenomena in a practicalcombustion process are dominated by turbulence, a fact thatcomplicates the effort to predict their characteristics. Althoughdirect numerical simulations (DNS) can resolve all the scalars of theNaviereStokes (NS) equations directly, because of its computa-tional complexity, DNS can only be applied to simple and smallgeometries and at lower Reynolds number. In situations where DNSis not feasible, turbulence models are used for approximatingturbulent fluxes, turbulent dispersion and turbulent-combustioninteractions. Several turbulence models have been developed;they are either based on the RANS or the LES approximations. InRANS-based models, dependent variables are decomposed intospace-time averaged components and fluctuations. The resultingReynolds fluxes are modeled by solving transport equations for theturbulent quantities, such as turbulent viscosity, turbulent kineticenergy, turbulent dissipation rate, etc., at all scales. On the otherhand, LES resolves the large eddies directly, while the impact of thesmall eddies are modeled. In terms of the fraction of the resolvedscales, LES falls between DNS and RANS [216]. CFD modeling of

ssurized 3 MWth pilot scale combustor under air-fired and oxy-fired modes. (a): mass. [80].

Page 50: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 205

turbulent combustion, especially of coal or other solid fuels, is stilllargely based on RANS.

As canbe seen inTable 12, the two-equation k� εmodel iswidelyused in oxy-coal combustion CFD simulations. These RANS simula-tions are reported to yield acceptable accuracywhen comparedwithexperimental data. Recently, Edge et al. [217] compared the perfor-mances of LES and RANS (RNG k� ε) models in their simulation ofair- and oxy-fired pulverized coal combustion in a 0.5 MWthcombustion test facility. Fig. 66 shows the comparison between thepredicted velocity and temperature fields using these twoapproaches under both air-fired (top) and oxy-fired (bottom)conditions. The contour maps show that LES predicts strongerrecirculation zones than RANS simulations in all cases. It also yieldsahotterflameedgebecauseof its ability to resolve the largeeddies ofhot gas. In general, similar improvements are observed when usingLES instead of RANS under air-fired and oxy-fired conditions. Thestudy also showed that simulations using a combination of LES anda gray gas radiation sub-model overpredicts the incident radiationwhile the combination of RANS and a spectral gas radiation sub-model yields better comparisons. This finding further accentuatesthe importance of other physical sub-models.

4.3. Radiation heat transfer

4.3.1. Models for radiative propertiesRadiation heat transfer is typically computed in CFD by solving

the radiative transfer equation (RTE). The RTE at position r.

in thedirection s

.is given by [218]:

dIlð r.

; s. Þ

ds¼ kln

2Ib;l ��kl þ ss;l

�Ilð r

.; s. Þ

þ ss;l4p

Z4p0

Ilð r.

; s. 0ÞFð s. $ s

. 0ÞdU0 (44)

where s is the path length, k is the absorption coefficient, ss isthe scattering coefficient, n is the refractive index, and I is theradiation intensity which is dependent on position r

.and direction

Fig. 66. (a) Axial velocity distribution from LES and RANS. Individual axial velocity legendsfrom LES and RANS. Temperature legend is shown in K. [217].

s.

. In this section, only the gaseous radiative heat transfer will bediscussed. The gas mixture is considered to be an absorbing andemitting medium without scattering. In this case, the RTE issimplified to:

dIlð r.

; s. Þ

ds¼ kln

2Ib;l � klIlð r.

; s. Þ (45)

Integrating the radiation intensity over the whole radiationspectrum yields:

dIð r. ; s. Þ

ds¼ an2

sT4

p� aincidentIð r

.; s. Þ (46)

The Planck-mean absorption coefficient, a, and the incidentmean absorption coefficient, aincident, are absorption coefficientsaveraged over the whole spectrumweighted by radiation intensity.Therefore, these properties are functions of gas composition,temperature and pressure; additionally, the incident meanabsorption coefficient, aincident, also depends on the incident radi-ation. An accurate calculation of the gas radiation property mustconsider the spectral absorption of CO2 and H2O, and also theoverlap between these radiating gas components. As discussed inSection 2.1.1, since direct prediction of the absorption coefficient iscomputationally complex, models for absorption properties, suchas the gray gas or the band models, will have to be employed in theradiation heat transfer model.

The most commonly used gray gas model in CFD is the weightedsum of gray gases (WSGG) model proposed by Smith et al. [219]. Inthis model, the gas is assumed to consist of a transparent gas andseveral gray gases without any wavelength dependence, and themodel parameters are based on the radiation properties of air-firedflue gases. In contrast, when using the band models, spectralcalculations are performed by dividing the entire wavelengthspectrum into several bands and assuming that the absorptioncharacteristics of each species remain either uniform or changesmoothly according to a given functional form over these bands[220]. One example of this type of radiation models is the Expo-nential Wide Band model (EWBM) proposed by Edwards and

are shown in m/s. (b) Velocity vectors are superimposed over temperature distribution

Page 51: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214206

Menard [221], which assumes exponential functions of the lineintensity around a band center. Band models are believed to beapplicable in oxy-fuel combustion because of their theoreticalsoundness in non-gray radiation. However the use of band modelsis usually computational expensive, and the model is typically usedto calibrate the gray gases model [222,223].

4.3.2. Modification of the gray-gas modelIn the WSGG model, the Planck-mean absorption coefficient of

the gas mixture over a path length, s, is determined by:

a ¼ �lnð1� εÞ=s (47)

where s is the radiation beam length, and ε, the gas emissivity. Thelatter is calculated from:

ε ¼Xi

aε;iðTÞð1� expð�kipisÞÞ (48)

where aε;i is the emissivity weighting factor for gray gas i, ki and pi

are the pressure absorption coefficient (1=m$atm) and partialpressure (atm) of the absorbing gas i, respectively.

The emissivity weighting factors used in the previous Eq. (48)are polynomial correlations that can be given as a function of thegas temperature:

aε;i ¼

Xj

bε;i;jT

j�1 (49)

The coefficients of the polynomial correlations, bε;i;j, and the

absorption coefficients of the gray gas, ki, are derived from exper-imental data of oil and methane stoichiometric combustion, inwhich the CO2 partial pressure is around 0.1 atm, and the partialpressure ratio of H2O and CO2 is in the range of 1e2, with a pathlength of less than 10 m [223]. It is worth noting that the CO2 andH2O partial pressures and their relative ratio in oxy-coal combus-tion are not within the applicable range.

Rehfeldt et al. [223] compared the predicted emissivity of theflue gases in both air-fuel and oxy-fuel combustion conditions at1000 K and 1500 K using both theWSGGmodel and EWBM, and theresults are shown in Fig. 67. Both models predict similar gasemissivities for the air-fuel conditions. However, the agreementbetween the two models is poor for oxy-fuel combustion condi-tions, especially at high CO2/H2O ratios.

a b

Fig. 67. Predicted total emissivity of gas mixture of carbon dioxide and water vapor at1500 K for: (a) conventional air combustion and (b) oxy-fuel combustion using theEWBM and the WSGG models. Values in parentheses show the molar fraction ofcarbon dioxide/water vapor [223].

Therefore, when using the WSGG model, the coefficients of thepolynomial correlations should be modified to reflect the condi-tions found in oxy-fuel combustion. Direct use of the gray gasesmodel for the radiative properties in oxy-fuel combustion maycause errors and should be avoided [218]. In order to adapt theWSGG model to oxy-fuel conditions, Johansson et al. [222] modi-fied its parameters by fitting the emissivity to a Statistical NarrowBand (SNB)model. The absorption coefficients are constant for eachgas, while the weights are calculated from temperature-dependentrelations. The two new parameter sets are applicable in oxy-firedconditions with dry or wet recycle, in a temperature range of500e2500 K and for path lengths between 0.01 and 60 m.

Likewise, Rehfeldt et al. [223] developed a four gray gases plusone clear gas approach. Instead of providing different parametersfor specified H2O mole fractions as done in reference [222], themolar ratio of H2O to CO2 (f) was considered as a variable in theemissivity weighting factor correlations:

aε;i ¼

Xj

"Xm

cε;i;j;mf

m�1

#Tj�1 (50)

where cε;i;j;m is the coefficient of the correlationwhich is derived

by curve fitting with the EWBM predictions in the range of600 Ke2400 K andmolar fractions of carbon dioxide in the range of0.3e0.9. A comparison between the predictions of radiative heatflux in a furnace using both the WSGG model with these newparameters and the EWBM shows good agreement.

4.4. Heterogeneous reactions

Oxy-char combustion characteristics have been reviewed inSection 2.6, which showed that the kinetics of char oxidation andgasification reactions are not significantly different from those inair-combustion. However, the mathematical model used in CFDsimulations should reflect the effects of the slower oxygen diffusionand the contributions of gasification reactions in a CO2-richatmosphere.

Chui et al. [201] modeled the char burning rate as being gov-erned by chemical kinetics and the external diffusion rate of oxygento the char surface, in a CFD simulation of oxy-coal combustion inthe CANMET vertical combustor research facility. The work onlyconsidered the char-O2 oxidation reactionwhich is implemented asa global carbon oxidation reaction:

CðsÞþO2/CO2 (51)

A similar model was used by Khare et al. [207] in a CFD simu-lation of oxy-coal combustion in the IHI vertical pilot-scale facility.Different mass diffusion-limited rate constants and kinetic-limitedrate parameters for air-coal and oxy-coal combustion derived fromDTF experimental results were used.

Taking the gasification reactions into account improves theprediction of char consumption rate, especially during the laterstages of oxy-combustion (also see 2.6.3). Therefore, in some CFDsimulations [189,224], the Boudouard reaction (Reaction III in Eq.(25)) was considered. In these studies, the product of carbonoxidation reaction is CO rather than CO2, (Reaction I in Eq. (23))which is reasonable for the combustion temperature range ofpulverized coal flames, as shown by Arthur’s experiments [225].Kuhr et al. [226] simulated char combustion in both O2/N2 and O2/CO2 atmospheres in a vertical once-through 20 kW furnace at theInstitute of Combustion and Power Plant Technology (IFK). The CFDstudy compared the performance of an intrinsic model based onSmith’s approach [161] against a surface reaction model. Fig. 68shows the experimental and numerical simulation results of the

Page 52: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

a b

Fig. 68. Measured and calculated species volume fractions for the lignite-char burningwith a wall temperature of 1300 �C (a) CFD predictions using intrinsic char oxidationmodel under 95% N2 and 5% O2 condition. (b) CFD predictions using surface reactionmodel under 95% CO2 and 5%O2 conditions [226].

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 207

O2 andCOmole fractionprofiles along the reactor. It should be notedthat the operating condition shown in Fig. 68 is 5% O2 with a walltemperatureof 1300 �C,which falls betweenRegionBandCof Fig. 32where the Boudouard reaction contributes to the char consumptionrate. As seen in Fig. 68(a), the intrinsic char oxidation model, whichonly considers the oxidation reaction, can predict the gas speciesdistribution well for the O2/N2 condition. However, when a surfacereactionmodel that consists of only theoxidation reaction is used forthe O2/CO2 condition, it underpredicts the char consumption rateand overpredicts the O2mole fraction as shown in Fig. 68(b). Addingthe Boudouard reaction increases the char consumption rate andimproves the O2 profile prediction. Additionally, since the partialpressure ofH2O is substantially higher underwet recycle conditions,most other CFD simulation studies [80,95,162,215,227] imple-mented both the char-CO2 and char-H2O gasification reactions(Reaction II in Eq. (24)) in the char consumption model, where thechar-H2O gasification reaction is considered to take place in parallelwith the oxidation reaction and Boudouard reaction.

The reaction rates of the three heterogeneous reactions dis-cussed above are determined by both surface reaction kinetics andexternal oxidizer diffusion. Based on the assumption that the globalreaction kinetics do not change in oxy-fuel combustion [148], thekinetic parameters are usually taken from air-coal combustionkinetics for similar ranks. On the other hand, in order to capture theeffect of slower diffusion rate under oxy-coal combustion condi-tions, the diffusion rates of the oxidizers in the boundary layer ofthe coal particles have to be modified based on the diffusion theory[161]. Doing so yields a lower diffusion rate constant under oxy-fired conditions (Ci ¼ 4:13� 10�12 in oxy-fuel vs.Ci ¼ 5:32� 10�12 in air-fuel, see Table 8). However, the change ofthis value was found to have negligible effects on the meantemperature distributionwithin the boiler with differences that areless than 5 K, as shown by Nikolopoulos et al. [224] in a full scale330 MWe PC boiler CFD modeling.

For computational simplicity, oxy-char combustion sub-modelsare usually simplified in CFD simulations. Important processes suchas the internal diffusion and reactions within the porous charparticles, the CO flame sheet around the char particles and theinteractions of gas species diffusion in the boundary layer areusually ignored. Therefore, advanced sub-models that are accurateand yet computationally inexpensive need to be developed for theoxy-char combustion.

4.5. Gas phase reactions

Gas phase reactions play an important role in coal combustionand gasification. Light gases, hydrocarbon gases (alkenes and alkyls)and tar are produced alongside CO2 and pyrolysis H2O during coaldevolatilization [121]. CO and H2 produced from the char particlegasification reactions are also burned in the free stream outside thecoal particle boundary layer for typical pulverized coal particle sizes[228]. In a CO2-rich environment, the global reaction rates might bedifferent from the conventional air-fuel combustion due to thechemical effects of CO2. Therefore, the modeling of gas phase reac-tions should be modified as well. This includes both the reactionmechanism and the turbulence-chemistry interaction model.

4.5.1. Reduced reaction mechanismsDetailed reaction mechanisms for hydrocarbon combustion

such as the detailed chemical kinetic model (DCKM) presented byGlarborg and Bentzen [188], and the GRI-Mech 3.0 used previouslyin this work have been shown to be valid under oxy-fuel combus-tion conditions. However, it is computationally expensive to applythese detailed reaction mechanisms in CFD modeling of coalcombustion. Alternatively, reduced hydrocarbon combustionmechanisms have been proposed.

Reduced reactionmechanisms andmodels used in CFD studies ofoxy-fuel combustion are summarized in Table 12. In early numericalsimulationsof oxy-coal combustion, volatilematter combustionwassimplified to a one-step global reaction. For instance, Wang et al.[32] simulated the oxy-coal combustion in the Battelle ColumbusLaboratory subscale combustor using an infinitely fast reaction rateand local chemical equilibrium. The simulation results showedsimilar trends to the experimental data for the temperature andmajor composition distributions. On the other hand, the modelfailed to predict the CO concentration because of the assumptions ofone-dimensional well-mixing and infinitely fast reactions.

Toporov et al. [162] proposed a simplified reaction mechanismfor volatile combustion in the CFD simulation of oxy-coalcombustion, which consists of volatile decomposition and irre-versible combustion of CO and H2:

CxHyOlSnNmþ�x2þn� l

2

�O2/xCOþy

2H2þnSO2þ

m2N2 (52)

COþ 0:5O2/CO2 (53)

H2 þ 0:5O2/H2O (54)

The simulation results showed agreement with experiments forthe temperature and O2 species. The CO concentration measure-ment was not available, and hence the accuracy of the CO predic-tion using this mechanism was not verified.

Multi-step reaction mechanisms, such as the two-step mecha-nism proposed by Westbrook and Dryer (WD) [229] and the four-step mechanism proposed by Jones and Lindstedt (JL) [230], weresuggested for the homogeneous reactions under conventional air-fired conditions, and have been successfully used in CFDmodeling of coal combustion and gasification [231,232]. Since theWD mechanism considers the reversible CO2 dissociation reaction,and the JL mechanism consists of the reversible water gas shiftreaction, they are expected to yield a more accurate CO concen-tration prediction than the irreversible reaction mechanisms.However, since the kinetics of these global reactions were deducedfrom air-combustion experimental data, Andersen and coworkers[233] concluded that these mechanisms cannot be expected towork as well under oxy-fuel conditions as they do for conventionalair-fuel combustion. Therefore, they further modified the two-step

Page 53: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214208

WD and four-step JL mechanisms for methane combustion by cal-ibrating their kinetics parameters with the ignition delay and thefinal product concentration results predicted by the DCKM reactionmechanism under oxy-fuel conditions. Both modified mechanismsare claimed to perform better than the original mechanisms for COand flame temperature predictions in the CFD modeling of oxy-propane combustion.

Coal combustion involves complicated hydrocarbon species,including both tar and gases that depend on the rank and propertyof the fuel, resulting in a large range of gas compositions. Althoughthe reduced mechanisms are reasonably well established fornatural gas and other gas fuels, and have been applied in air-coalcombustion, their accuracies in oxy-coal combustion simulationremain a subject of investigation.

4.5.2. Turbulence-chemistry interactionsFor laminar flows, the reaction rates can be computed using the

laminar finite-rate model. However, in the case of turbulent flames,turbulent fluctuations becomes important and chemistry-turbulence interactions must be modeled while considering theeffects of turbulent fluctuations on the source term of the speciesequation. The description of the interaction between turbulenceand chemistry, therefore, becomes another concern in accuratemodeling the gas phase reaction under oxy-fuel conditions.

The eddy dissipation model (EDM) [234] was used in somenumerical studies of oxy-coal combustion [95,201] in which thechemical reaction is governed by the large-eddy mixing time scalebased on the eddy break-up (EBU)model proposed by Spaling [235].As a further development, the finite rate/eddy dissipation modelwas proposed, in which the reaction rate is taken to be the lower ofthe Arrhenius reaction rate and eddy dissipation rate. In the flamezone, the eddy dissipation rate is generally smaller than theArrhenius rate, and reactions are limited by the extent of mixing.Based on the assumption that the reaction rate is limited by both themixing of the reactants and the heating of the reactants through themixing with hot reaction products, the net rate of production ofspecies i due to reaction r, Ri;r , is given by the smaller of the reactantmixing rate (first term) and product mixing rate (second term):

Ri;r ¼min

n0i;rMw;iAr

ε

kminR

�YR

n0R;rMw;R

�; n0i;rMw;iABr

ε

k

PPYPPN

j n00j;rMw;j

!

(55)

Where A and B are model constants n0i;r and n00j;r are the stoi-chiometric coefficient for reactant i and product j in reaction r,Mw isthe molecular weight, YR and YP are mass fraction of any reactantand product species, respectively. Note that in oxy-fuel combustion,the CO2 and H2O partial pressures in the burner streams are alreadyhigh before mixing with hot product gases. Therefore, using thesame model constants A and B for air-fuel combustion may over-estimate the second term, and thus over-estimate the reactionrate of oxy-fuel combustion.

Another drawback of eddy dissipation based models is that theycannot deal with multi-step chemical mechanisms and reversiblereactions, such as CO2 dissociation. The reason is that multi-stepmechanisms are based on reaction rates which occur at differenttime scales, whereas in these models, every reaction has the sameturbulent rate, and therefore, the same reaction rate [216]. CFDmodeling of an oxy-natural gas flame has shown that the mixed-is-burned nature of the eddy dissipation model results in a failure topredict CO accurately [236].

The eddy dissipation concept (EDC) model was developed as anextension to the eddy dissipation model to incorporate multi-stepchemical mechanisms in turbulent flows [237], where the reac-tions are assumed to occur in a turbulence-characterized fine-scale

reactor governed by Arrhenius rates. Multi-step reaction mecha-nisms have been implemented in this model for air-methanecombustion [238,239] as well as air-coal combustion [240], andare believed to be more accurate than the eddy dissipation basedmodels. Recently, Vascellari and Cau [227] applied this model witha modified JL mechanism which is taken from air-fuel combustion,to simulate oxy-coal combustion in the IFRF 2.4 MWth furnace.However they did not show a comparison with experimental data.Similarly, Muller et al. [215] also used the same approach for the gasphase reactions in their simulation of oxy-coal combustion in theIFK 0.5 MWth test facility, and the CO prediction agrees well withtheir experimental data.

4.6. Summary

The sub-models for turbulence, reaction and radiation inconventional air-coal combustion have been modified to take intoaccount the different characteristics of oxy-coal combustion. Theeffects of the physical properties of the CO2-rich stream oncombustion have been investigated using CFD approaches, offeringinsight into oxy-coal combustion. However, the development andvalidation of more accurate sub-models are still needed. Extensiveapplications of CFD are expected in the scale up and advanceddesign of oxy-coal combustion facilities.

5. Research need

Oxy-coal combustion is a promising novel technology, and it isbeing implemented at scale. New challenges have been identifiedas this technology is developed, from the system level, such assystem integration and optimization, to the fundamental levels,such as stabilization of oxy-fuel flames. Based on this review, futureresearch needs are suggested below.

5.1. Oxy-coal system design

� CO2 purity standard for sequestration: Up to the time of thisreview, there is still no consensus on the required CO2 purityfor sequestration. The standard should consider both safetyrequirements and the operating costs of capture, trans-portation and site monitoring. The standard will determine theefficiency and cost of electricity from oxy-coal combustiontechnology. An overly high cost to meet the standard maybecome a hindrance of its application (for instance, the strin-gent oxygen content required for EOR).

� System integration and optimization: Before oxy-coal combus-tion systems can be employed, the system componentsmust beintegrated, and the operating conditions optimized for the bestenergy efficiency. Thus, detailed process designs of a large-scale oxy-coal power plant are required for both retrofits andnew plants. Since the operating conditions, such as the recycleratio, are coal dependent, guidelines of system designs andoperating conditions for different coal ranks are also required.

� New oxy-coal combustion processes for carbon capture: As dis-cussed previously, most of the current studies have focused onair-like oxy-coal combustion and matching the combustionand heat transfer characteristics with those in air-combustion.Meanwhile, other oxy-combustion regimes mentioned inSection 1.2, such as full oxy-combustion and MILD oxy-combustion processes, may lead to higher efficiency andlower cost and thus, should continue to be explored. Similarly,pressurized oxy-coal systems are promising, but they are still ina nascent stage. Pilot scale tests for these concepts are essentialto make them strong contenders to the more established air-like atmospheric oxy-coal system.

Page 54: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 209

5.2. Fundamentals of oxy-coal combustion

Considerable knowledge of the fundamentals of heat and masstransfer, combustion processes and pollutant formation under CO2-rich conditions has been acquired through extensive experimentsand modeling efforts in recent years. However, many fundamentalissues remain uninvestigated:

� Oxy-combustion characteristics of different coal types: Due to thedistinct physical properties and chemical structures, studieshave shown various characteristics for coal of different ranksduring devolatilization, volatile combustion, char burning, andpollutant formation. Their characteristics under oxy-fuelcombustion should be further investigated as an integral partof oxy-burner design.

� Models for sub-processes: Models for devolatilization, charformation, agglomeration and coal group combustion underthe oxy-combustion conditions are mostly based on aircombustion studies. For instance devolatilization measure-ments are usually taken in inert medium, while in oxy-combustion, pyrolysis occurs in a CO2-rich environment inwhich some chemical interactions might occur. Similarly, charstructure and consumption submodels are mostly formulatedfor combustion in air, with a typical oxygen concentration of21%. In oxy-combustion, char oxidation sub-models may needto be modified depending on the temperature, oxygenconcentration and the presence of a higher fraction of steam.

� Scaling characteristics of oxy-combustion: Given the complexityof combustion processes and the number of nonlinearlyinteracting processes, developing similarity and scaling modelsis usually challenging. Extending such models acquired fromexperience with traditional air-combustion to oxy-combustionis necessary in order to establish design guidelines and rulesfor burner and combustor sizing.

� Pressure effect: When operating under elevated pressure, thegas phase flowfield and coal particle residence time maychange significantly. The effect of pressure on the characteris-tics of mass and heat transfer, char and gas phase combustionkinetics and dynamics, etc, are also still relatively unknown.Research on the characteristics of oxy-combustion at elevatedpressures is needed.

5.3. CFD modeling of oxy-coal combustion

Although CFD approaches have been used in some studies tobetter understand the flowfield and combustion processes in oxy-coal combustion, several problems remain that need to beresolved to achieve a higher predictive accuracy of combustioncharacteristics in a CO2-rich environment. Models of some sub-processes are not yet available or verified in oxy-coal combustion,such as:

� Oxy-char combustion sub-models: The models should take intoaccount the effect of physical (heat capacity and mass transfer)and chemical (its interaction with heterogeneous andhomogenous reactions) properties of CO2 and be able to predictthe burning rate in the different operating condition regions(see Section 2.6).

� Gas phase reaction mechanisms: High-fidelity reduced gascombustion mechanism should capture the chemical effect ofCO2 in oxy-combustion, and provide accurate predictions ofminor species and pollutant formations.

� Radiation models: Radiation heat transfer plays a major role inthe furnace, and it also governs the energy equation in

combustion. Better turbulence-radiation and gas emittance/absorptance sub-models are needed to improve the simulationof the temperature field.

� Turbulence models: While swirling flows and their impact onmixing via the establishment of inner recirculation zones nearburners and injectors are important in all combustors, theytake on a more significant role in the case of oxy-combustion.Better predictive models for rotating and swirling flows arethus needed, especially for the more economical approaches,such as RANS.

5.4. Large scale demonstration

Pilot and industrial scale demonstration of the oxy-coalcombustion is imperative to verify the observations and theoriesdeveloped from lab and bench scale studies, and provide valuableoperating experience for future scale-ups.

� Transition between operating modes: Considering the uncer-tainty of greenhouse gas emission regulations, future powerplant may need to be designed to operate under both air- andoxy-fired conditions. The capability of operating under bothmodes is also one of the key advantages because the air-firingoperation is needed during start up/shut down, peak hourdemands, and auxiliary equipment (such as ASU and CO2

compression units) failures. Currently, the transition betweenthese operating modes is being tested in pilot scale facilities ofVattenfall, Alstom and Doosan Babcock and shows little diffi-culties [241e244]. However, the steam streams and energyflows are heavily coupled in demonstration scale oxyfuelcombustion system, and issues such as the control of the sub-systems coupling during start-up still need to be addressed.

� Advanced oxy-coal burner and combustor design and perfor-mance test: As previously discussed in Section 3, efforts arebeing made to understand the fundamentals of flame desta-bilization in oxy-coal combustion, as well as the principles ofretrofitting and designing the oxy-coal burners. Howevermuchmore remains to be done for advanced oxy-fuel burner design,its carbon conversion and pollutant control performance, aswell as its scale-up to full-scale power plants. Following theonsite tests of single oxy-burner performance, the demon-stration of larger burners and multi-burner configurationsshould also be carried out.

� Material and corrosion: Since the SO2 and NOx concentrationcan be higher in the oxy-fuel mode, high temperature corro-sion of various alloy materials should be tested at typicallocations in the flue gas path to serve as guide for choosing thematerial of heat exchangers. Low temperature corrosion test inthe flue gas duct is also needed to set up standards for the acidgases in recycled flue gas [243].

� Fate of ash and slag: While the emissions of pollutants like NOxand SO2 are widely studied in oxy-coal combustion, there isa lack of literature on the behaviors of ash and slag in thisenvironment. Little is known about the phase changes, coag-ulation, size distribution and element partitions of the fly ash ina high CO2 environment. Particulate matter control deviceshave to be adapted based on these studies. In the case ofslagging oxy-fuel combustion, the mechanisms of slag layerformation and its interactionwith char particles are also poorlyunderstood.

This paper reviewed the state-of-the-art progresses of oxy-coalcombustion on aspects of its regime characterization, fundamen-tals, stabilization and CFD modeling approaches. The discussionsfocus on the different characteristics between oxy-fuel combustion

Page 55: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214210

and the conventional air-fuel combustion, and reveal insight on thephysical and chemical effects in different combustionmedia. Due toits complexity, the combustion technology is still heavily depen-dent on experiments and operational experience, especially in itsapplication in coal-fired power plants. Therefore, successes in pilot-and demonstration-scale tests such as the Vattenfall project shouldprovide practical knowledge and experience for the operation, heattransfer, combustion dynamics and stability, as well as pollutantformation and control.

Acknowledgment

The authors appreciate Dr. Marco Gazzino of ENEL Ingegneria eInnovazione S.p.A for the helpful discussions on pressurized oxy-fuel coal combustion, Professor Janos M. Beer of MIT for hisadvice on the fundamentals of coal combustion aerodynamics, andMr. Simcha Singer for the discussions on aspects of coal charcombustion models.

References

[1] IPCC. Contribution of working groups I, II and III to the fourth assessmentreport of the intergovernmental panel on climate change. IntergovernmentalPanel on Climate Change; 2007.

[2] Ghoniem AF. Needs, resources and climate change: clean and efficientconversion technologies. Progress in Energy and Combustion Science 2011;37:15e51.

[3] Pacala S, Socolow R. Stabilization wedges: solving the climate problem forthe next 50 years with current technologies. Science 2004;305:968e72.

[4] BP. Statistical review of world energy 2010; 2010.[5] Annual energy review 2009. U.S. Energy Information Administration; 2010.[6] Wall TF. Combustion processes for carbon capture. Proceedings of the

Combustion Institute 2007;31:31e47.[7] Wang M, Lawal A, Stephenson P, Sidders J, Ramshaw C. Post-combustion CO2

capture with chemical absorption: a state-of-the-art review. ChemicalEngineering Research and Design, in press, Corrected Proof.

[8] Blamey J, Anthony EJ, Wang J, Fennell PS. The calcium looping cycle for large-scale CO2 capture. Progress in Energy and Combustion Science 2010;36:260e79.

[9] Thiruvenkatachari R, Su S, An H, Yu XX. Post combustion CO2 capture bycarbon fibre monolithic adsorbents. Progress in Energy and CombustionScience 2009;35:438e55.

[10] Merkel TC, Lin H, Wei X, Baker R. Power plant post-combustion carbondioxide capture: an opportunity for membranes. Journal of MembraneScience 2010;359:126e39.

[11] Lyngfelt A, Leckner B, Mattisson T. A fluidized-bed combustion process withinherent CO2 separation; application of chemical-looping combustion.Chemical Engineering Science 2001;56:3101e13.

[12] Hossain MM, de Lasa HI. Chemical-looping combustion (CLC) for inherentCO2 separationsea review. Chemical Engineering Science 2008;63:4433e51.

[13] Fan L- S, Wiley InterScience (Online service). Chemical looping systems forfossil energy conversions. Hoboken, NJ: Wiley-AIChE; 2010.

[14] The future of coal: options for a carbon-constrained world. MIT; 2007.[15] Haslbeck JL, Capicotto PJ, Juehn NJ, Lewis E, Rutkowski MD, Woods MC, et al.

Pulverized coal oxycombustion power plants. In: Bituminous coal to elec-tricity, Vol. 1. Washington D.C.: DOE/NETL-2007/1291; 2007.

[16] Kanniche M, Gros-Bonnivard R, Jaud P, Valle-Marcos J, Amann J-M,Bouallou C. Pre-combustion, post-combustion and oxy-combustion inthermal power plant for CO2 capture. Applied Thermal Engineering 2010;30:53e62.

[17] EIA-860 annual electric generator report. U.S. Energy Information Adminis-tration; 2010.

[18] Shuster E. Tracking new coal-fired power plants. National Energy TechnologyLaboratory; 2011.

[19] DOE US. Secretary Chu Announces FutureGen 2.0, http://www.energy.gov/9309.htm; 2010, Aug 5th.

[20] Toftegaard MB, Brix J, Jensen PA, Glarborg P, Jensen AD. Oxy-fuel combustionof solid fuels. Progress in Energy and Combustion Science 2010;36:581e625.

[21] Buhre BJP, Elliott LK, Sheng CD, Gupta RP, Wall TF. Oxy-fuel combustiontechnology for coal-fired power generation. Progress in Energy andCombustion Science 2005;31:283e307.

[22] Wall T, Liu Y, Spero C, Elliott L, Khare S, Rathnam R, et al. An overview onoxyfuel coal combustion e State of the art research and technology devel-opment. Chemical Engineering Research and Design 2009;87:1003e16.

[23] Wall T, Yu J. Coal-fired oxyfuel technology status and progress to deploy-ment. In: The 34th international technical conference on clean coal & fuelsystems. Clearwater: Florida, USA; 2009.

[24] Baukal CE. Oxygen-enhanced combustion. Boca Raton, Fla: CRC Press; 1998.

[25] Santos S, Haines M. Oxy-fuel combustion application for coal fired powerplant. IEA Greenhouse Gas R&D Programme. Cottbus, Germany; 2005.

[26] Tsuji H. High temperature air combustion: from energy conservation topollution reduction. Boca Raton, FL: CRC Press; 2003.

[27] Cavaliere A, de Joannon M. Mild combustion. Progress in Energy andCombustion Science 2004;30:329e66.

[28] Horn FL, Steinberg M. Control of carbon dioxide emissions from a powerplant (and use in enhanced oil recovery). Fuel 1982;61:415e22.

[29] Steinberg M. History of CO2 greenhouse gas mitigation technologies. EnergyConversion and Management 1992;33:311e5.

[30] Abraham B, Asbury J, Lynch E, Teotia A. Coal-oxygen process provides carbondioxide for enhanced recovery. Oil Gas Journal 1982;80:68e70.

[31] Abele AR, Kindt GS, Clark WD, Payne R, Chen SL. An experimental program totest the feasibility of obtaining normal performance from combustors usingoxygen and recycled gas instead of air. Argonne National Lab Report ANL/CNSV-TM-204; 1988.

[32] Wang CS, Berry GF, Chang KC, Wolsky AM. Combustion of pulverized coalusing waste carbon dioxide and oxygen. Combustion and Flame 1988;72:301e10.

[33] Payne R, Chen SL, Wolsky AM, Richter WF. CO2 recovery via coal combustionin mixtures of oxygen and recycled flue gas. Combustion Science andTechnology 1989;67:1e16.

[34] Zheng L. Oxy-fuel combustion for power generation and carbon dioxide(CO2) capture. Philadelphia: Woodhead Publishing Limited; 2011.

[35] Herzog H. CCS project database, http://sequestration.mit.edu/tools/projects/index.html; 2011.

[36] Zanganeh KE, Salvador C, Mitrovic M, Shafeen A. 3rd generation oxy-fuelcombustion systems. In: 3rd IEAGHG international oxy-combustionnetwork. Yokohama, Japan; 2008.

[37] Schaffel N, Mancini M, Szlek A, Weber R. Mathematical modeling of MILDcombustion of pulverized coal. Combustion and Flame 2009;156:1771e84.

[38] Zhang H, Yue G, Lu J, Jia Z, Mao J, Fujimori T, et al. Development of hightemperature air combustion technology in pulverized fossil fuel fired boilers.Proceedings of the Combustion Institute 2007;31:2779e85.

[39] Zhang H, Cui K, Li Z, Wang T, Liu Q, Yang H, et al. Development of oxygen-enriched high temperature air combustion burner for PC boilers. In: The34th international technical conference on coal utilization and fuel systems.Clearwater, Florida, USA; 2009.

[40] Stadler H, Ristic D, Forster M, Schuster A, Kneer R, Scheffknecht G. NOx-emissions from flameless coal combustion in air, Ar/O2 and CO2/O2.Proceedings of the Combustion Institute 2009;32:3131e8.

[41] Weber R, Smart JP, Kamp Wv. On the (MILD) combustion of gaseous, liquid,and solid fuels in high temperature preheated air. Proceedings of theCombustion Institute 2005;30:2623e9.

[42] Seepana S, Jayanti S. A new stable operating regime for oxyfuel combustion.In: ASME international mechanical engineering congress and exposition.Boston, MA; 2008: 435e444.

[43] Calchetti G, Nardo AD, Mongibello G, Mongiello C. Mild combustion simu-lation of coal water slurry. In: Italian section of the combustion institute 30thmeeting on combustion. Ischia, Italy; 2007.

[44] Nardo AD, Calchetti G, Mongibello G, Caprariis Bd, Rufoloni M. Coal waterslurry oxy-combustion CFD analysis. In: 1st International conference onsustainable fossil fuels for future energy. Rome, Italy; 2009.

[45] Dunn-Rankin D. Lean combustion: technology and control. Amsterdam/Boston: Academic Press; 2008.

[46] Seepana S, Jayanti S. Steam-moderated oxy-fuel combustion. EnergyConversion and Management 2010;51:1981e8.

[47] Hong J, Chaudhry G, Brisson JG, Field R, Gazzino M, Ghoniem AF. Analysis ofoxy-fuel combustion power cycle utilizing a pressurized coal combustor.Energy 2009;34:1332e40.

[48] Andersen RE, MacAdam S, Viteri F, Davies DO, Downs JP, Paliszewski A.Adapting gas turbines to zero emission oxy-fuel power plants. In: Proceed-ings of ASME Turbo Expo 2008: power for land, sea and air. Berlin, Germany;2008.

[49] Zheng L, Pomalis R, Clements B, Herage T. Optimization of a high pressureoxy-fuel combustion process for power generation and CO2 capture. In: The35th international technical conference on clean coal & fuel systems.Clearwater, FL; 2010.

[50] Fassbender AG, Henry RS, Tao L. Analysis of ThermoEnergy integrated powersystem for new power facilities using Alaskan and reference coal types. AEAReport, AEA Grant Number - AEA 07-014; 2009.

[51] Hong JMS. Thesis: techno-economic analysis of pressurized oxy-fuelcombustion power cycle for CO2 capture. Cambridge: Massachusetts Insti-tute of Technology; 2009.

[52] Hong J, Field R, Gazzino M, Ghoniem AF. Operating pressure dependence ofthe pressurized oxy-fuel combustion power cycle. Energy 2010;35:5391e9.

[53] Yang T, Hunt P, Lisauskas R, Ballas EN, Vitalis B. Pressurized OxycombustionCarbon Capture Power System. In: The 35th international technical confer-ence on clean coal & fuel systems. Clearwater, FL; 2010.

[54] BERR carbon abatement technologies programme: future CO2 capturetechnology options for the Canadian market. AEA Energy & Environment,COAL R309 BERR/Pub URN 07/1251; 2007.

[55] Darde A, Prabhakar R, Tranier J-P, Perrin N. Air separation and flue gascompression and purification units for oxy-coal combustion systems. EnergyProcedia 2009;1:527e34.

Page 56: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 211

[56] Andersson K, Johnsson F. Process evaluation of an 865 MWe lignite fired O2/CO2 power plant. Energy Conversion and Management 2006;47:3487e98.

[57] Okawa M, Kimura N, Kiga T, Takano S, Arai K, Kato M. Trial design for a CO2recovery power plant by burning pulverized coal in O2/CO2. EnergyConversion and Management 1997;38:S123e7.

[58] Varagani R, Chatel-Pelage F, Pranda P, Rostam-Abadi M, Lu Y, Bose A.Performance simulation and cost assessment of oxy-combustion process forCO2 capture from coal-fired power plants. In: The 4th annual conference oncarbon sequestration. Alexandria, VA; 2005.

[59] Dyer PN, Richards RE, Russek SL, Taylor DM. Ion transport membrane tech-nology for oxygen separation and syngas production. Solid State Ionics 2000;134:21e33.

[60] Moghtaderi B. Application of chemical looping concept for air separation athigh temperatures. Energy & Fuels 2009;24:190e8.

[61] Nakayama S, Noguchi Y, Kiga T, Miyamae S, Maeda U, Kawai M, et al.Pulverized coal combustion in O2/CO2 mixtures on a power plant for CO2recovery. Energy Conversion and Management 1992;33:379e86.

[62] Dillon DJ, White V, Allam RJ, Wall RA, Gibbins J. Oxy-combustion processesfor CO2 capture from power plant. Engineering investigation report IEAgreenhouse gas R&D research and development programme; 2005.

[63] Badzioch S, Hawksley PGW. Kinetics of Thermal decomposition of pulverizedcoal particles. Industrial & Engineering Chemistry Process Design andDevelopment 1970;9:521e30.

[64] Gazzino M, Benelli G. Pressurised oxy-coal combustion Rankine-cycle forfuture zero emission power plants: process design and energy analysis. In:Proceedings of ASME ES2008. Jacksonville, Florida, USA; 2008.

[65] CANMET. Clean fossil fuels: carbon capture & storage, high pressure oxy-fuel(HiPrOx), http://canmetenergy-canmetenergie.nrcan-rncan.gc.ca/eng/clean_fossils_fuels/carbon_capture_storage/hiprox.html; 2009.

[66] Zheng L, Pomalis R, Clements B. Technical and economic feasibility study ofa pressurized oxy-fuel approach to carbon capture: Part 1-Technical feasi-bility study and comparison of the ThermoEnergy Integrated Power System(TIPS) with a conventional power plant and other carbon capture processes.CANMET Energy Research Centre, Natural Resources Canada; 2007.

[67] Rossetti E, Malavasi M. Method and plant for the treatment of materials inparticular waste materials and refuse; 2004.

[68] Malavasi M, Rossetti E. High-efficiency combustors with reduced environ-mental impact and processes for power generation derivable Therefrom;2005.

[69] Malavasi M, Di Salvia G, Rossetti E. Combustion process; 2009.[70] White V, Torrente-Murciano L, Sturgeon D, Chadwick D. Purification of

oxyfuel-derived CO2. Energy Procedia 2009;1:399e406.[71] Fogash K, White V. Oxycoal combustion: opportunities and challenges -

purification of oxyfuel-derived CO2. In: AIChE 2007 Annual Meeting. SaltLake City, UT; 2007.

[72] Iloeje C, Field R, Gazzino M, Ghoniem A. Process modeling and analysis ofCO2 purification for pressurized oxy-coal combustion. In: The 35th Inter-national Technical Conference on Clean Coal and Fuel Systems. Clearwater,FL; 2010.

[73] Mathieu P, Nihart R. Zero-emission MATIANT Cycle. Journal of Engineeringfor Gas Turbines and Power 1999;121:116e20.

[74] Jackson AJB, Audus H, Singh R. Gas turbine engine configurations for powergeneration cycles having CO2 sequestration. Proceedings of the Institution ofMechanical Engineers, Part A: Journal of Power and Energy 2004;218:1e13.

[75] Richards GA, Casleton KH, Chorpening BT. CO2 and H2O diluted oxy-fuelcombustion for zero-emission power. Proceedings of the Institute ofMechanical Engineerings, Part A: Journal of Power and Energy 2005;219:121e6.

[76] Williams TC, Shaddix CR, Schefer RW. Effect of syngas composition and CO2-diluted oxygen on performance of a premixed swirl-stabilized combustor.Combustion Science and Technology 2008;180:64e88.

[77] Pomalis R, Zheng L, Clements B. ThermoEnergy integrated power systemeconomics. In: 32nd international technical conference on coal utilization &fuel systems. Clearwater, FL; 2007.

[78] Fassbender AG. Pressurized oxy-fuel combustion for multi-pollutant capture.In: The 30th international technical conference on coal utilization & fuelsystems. Clearwater, FL; 2005.

[79] Henry RS. Modeling of pressurized oxy-fuel for recovery of latent heat andcarbon capture using usibelli coal. In: 32nd international technical confer-ence on coal utilization & fuel systems. Clearwater, FL; 2007.

[80] Chen L, Gazzino M, Ghoniem AF. Characteristics of pressurized oxy-coalcombustion under increasing swirl number. In: The 35th internationaltechnical conference on clean coal & fuel systems. Clearwater, FL; 2010.

[81] Herwig H, Warmeubertragung A-Z. Systematische und ausführliche Erläu-terungen wichtiger Größen und Konzepte. Springer-Verlag Berlin and Hei-delberg GmbH & Co. KG; 2000.

[82] Edwards DK. Radiation heat transfer notes. Washington D.C.: Hemisphere;1981.

[83] Gupta RP, Wall TF. The optical properties of fly ash in coal fired furnaces.Combustion and Flame 1985;61:145e51.

[84] Wall TF, Stewart IM. The measurement and prediction of solids- andsoot-absorption coefficients in the flame region of an industrial P.F. chamber.Symposium (International) on Combustion 1973;14:689e97.

[85] Andersson K, Johnsson F. Flame and radiation characteristics of gas-fired O2/CO2 combustion. Fuel 2007;86:656e68.

[86] Andersson K, Johansson R, Hjartstam S, Johnsson F, Leckner B. Radiationintensity of lignite-fired oxy-fuel flames. Experimental Thermal and FluidScience 2008;33:67e76.

[87] Andersson K, Johansson R, Johnsson F, Leckner B. Radiation intensity ofpropane-fired oxy-fuel flames: implications for soot formation. Energy &Fuels 2008;22:1535e41.

[88] Hjartstam S, Andersson K, Johnsson F, Leckner B. Combustion characteristicsof lignite-fired oxy-fuel flames. Fuel 2009;88:2216e24.

[89] Woycenko D, Ikeda I, van de Kamp W. Combustion of pulverized coal ina mixture of oxygen and recycled flue gas. Technical Report Doc F98/Y/1.International Flame Research Foundation; 1994.

[90] Woycenko D, van de Kamp W, Roberts P. Combustion of pulverized coal ina mixture of oxygen and recycled flue gas. Technical Report Doc F98/y/2.International Flame Research Foundation; 1995.

[91] Smart JP, O’Nions P, Riley GS. Radiation and convective heat transfer, andburnout in oxy-coal combustion. Fuel 2010;89:2468e76.

[92] Croiset E, Thambimuthu K, Palmer A. Coal combustion in O2/CO2 mixturescompared with air. The Canadian Journal of Chemical Engineering 2000;78:402e7.

[93] Croiset E, Thambimuthu KV. NOx and SO2 emissions from O2/CO2 recyclecoal combustion. Fuel 2001;80:2117e21.

[94] Tan Y, Croiset E, Douglas MA, Thambimuthu KV. Combustion characteristicsof coal in a mixture of oxygen and recycled flue gas. Fuel 2006;85:507e12.

[95] Nozaki T, S-i Takano, Kiga T, Omata K, Kimura N. Analysis of the flameformed during oxidation of pulverized coal by an O2eCO2 mixture. Energy1997;22:199e205.

[96] Smart JP, Patel R, Riley GS. Oxy-fuel combustion of coal and biomass, theeffect on radiative and convective heat transfer and burnout. Combustionand Flame 2010;157:2230e40.

[97] Beer JM, Chomiak J, Smoot LD. Fluid dynamics of coal combustion: a review.Progress in Energy and Combustion Science 1984;10:177e208.

[98] Cabrejos FJ, Klinzing GE. Pickup and saltation mechanisms of solid particlesin horizontal pneumatic transport. Powder Technology 1994;79:173e86.

[99] NIST Chemistry WebBook: thermophysical properties of fluid systems,http://webbook.nist.gov/chemistry/; 2009.

[100] Villareal JA, Klinzing GE. Pickup velocities under higher pressure conditions.Powder Technology 1994;80:179e82.

[101] Klinzing GE. Pneumatic conveying of solids: a theoretical and practicalapproach. 2nd ed. London/New York: Chapman and Hall; 1997.

[102] NETL. Coal storage and feed preparation, http://www.netl.doe.gov/technologies/coalpower/gasification/gasifipedia/5-support/5-1_storage.html; 2011.

[103] Beer J. Clean combustion of coal, research and applications, an overview ofrecent development in the USA. In: Institute of energy: 51st Melchett lecture.Cafe Royal, London; 1985.

[104] Arthur DL. Investigation of low-rank-liquid carbon dioxide slurries. EPRIReport; 1986.

[105] Neville AHH. Coal gasification research, development and demonstrationneed and opportunities. In: Gasification technologies conference. San Fran-cisco, CA; 2001.

[106] Beer JM, Chigier NA. Combustion aerodynamics. Malabar, Fla: Krieger; 1983.[107] Sojka PE, Lefebvre AH. A novel method of atomizing coal-water slurry fuels:

final report for Grant No. DE-FG22-87PC79931 US DOE. West Lafayette, IN:Thermal Sciences and Propulsion Center, School of Mechanical Engineering,Purdue University; 1990. 1e116.

[108] Yan W, Zhao G-Y, Chen G-J, Guo T- M. Interfacial tension of(methane þ nitrogen) þ water and (carbon dioxide þ nitrogen) þ watersystems. Journal of Chemical & Engineering Data 2001;46:1544e8.

[109] Turns SR. An introduction to combustion: concepts and applications. 2nd ed.Boston: McGraw-Hill; 2000.

[110] Cen K, Yao Q, Cao X, Zhao X, Huang Z, Zhou J, et al. Theory and application ofcombustion, flow, heat transfer and gasification of coal slurry. Hangzhou:Zhejiang University Press; 1997.

[111] Mills AF. Heat transfer. Homewood, IL: Irwin; 1992.[112] Okasha F, Miccio M. Prediction of coal-water slurry dispersion in a fluidized

bed combustor. Symposium (International) on Combustion 1996;26:3277e85.

[113] Ranz WE, Marshall WR. Evaporation from drops, part I. Chemical engineeringprogress, vol. 48; 1952.

[114] Suuberg EM, Peters WA, Howard JB. Product compositions in rapid hydro-pyrolysis of coal. Fuel 1980;59:405e12.

[115] Anthony DB, Howard JB, Hottel HC, Meissner HP. Rapid devolatilization andhydrogasification of bituminous coal. Fuel 1976;55:121e8.

[116] Bautista JR, Russel WB, Saville DA. Time-resolved pyrolysis product distri-butions of softening coals. Industrial and Engineering Chemistry Funda-mentals 1986;25:536e44.

[117] Serio MA, Hamblen DG, Markham JR, Solomon PR. Kinetics of volatileproduct evolution in coal pyrolysis: experiment and theory. Energy Fuels1987;1:138e52.

[118] Gibbins-Matham J, Kandiyoti R. Coal pyrolysis yields from fast and slowheating in a wire-mesh apparatus with a gas sweep. Energy Fuels 1988;2:505e11.

[119] Ko GH, Sanchez DM, Peters WA, Howard JB. Correlations for effects of coaltype and pressure on tar yields from rapid devolatilization. Symposium(International) on Combustion 1989;22:115e24.

Page 57: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214212

[120] Li C- Z. Advances in the science of Victorian brown coal. Amsterdam/London:Elsevier; 2004.

[121] Chen L, Zeng C, Guo X, Mao Y, Zhang Y, Zhang X, et al. Gas evolution kineticsof two coal samples during rapid pyrolysis. Fuel Processing Technology2010;91.

[122] Donald B, Anthony JBH. Coal devolatilization and hydrogastification. AIChEJournal 1976;22:625e56.

[123] Rathnam RK, Elliott LK, Wall TF, Liu Y, Moghtaderi B. Differences in reactivityof pulverised coal in air (O2/N2) and oxy-fuel (O2/CO2) conditions. FuelProcessing Technology 2009;90:797e802.

[124] Al-Makhadmeh L, Maier J, Scheffknecht G. Coal pyrolysis and charcombustion under oxy-fuel conditions. In: The 34th international technicalconference on coal utilization & fuel systems. Clearwater, FL; 2009.

[125] Borrego AG, Alvarez D. Comparison of chars obtained under oxy-fuel andconventional pulverized coal combustion atmospheres. Energy & Fuels 2007;21:3171e9.

[126] Brix J, Jensen PA, Jensen AD. Coal devolatilization and char conversion undersuspension fired conditions in O2/N2 and O2/CO2 atmospheres. Fuel 2010;89:3373e80.

[127] Howard JB, Essenhigh RH. Pyrolysis of coal particles in pulverized fuelflames. Industrial & Engineering Chemistry Process Design and Development1967;6:74e84.

[128] Howard JB, Essenhigh RH. Mechanism of solid-partical combustion withsimultaneous gas-phase volatiles combustion. Symposium (International) onCombustion 1967;11:399e408.

[129] Essenhigh RH, Misra MK, Shaw DW. Ignition of coal particles: a review.Combustion and Flame 1989;77:3e30.

[130] Ponzio A, Senthoorselvan S, Yang W, Blasiak W, Eriksson O. Ignition of singlecoal particles in high-temperature oxidizers with various oxygen concen-trations. Fuel 2008;87:974e87.

[131] Levendis YA, Joshi K, Khatami R, Sarofim AF. Combustion behavior in air ofsingle particles from three different coal ranks and from sugarcane bagasse.Combustion and Flame 2011;158:452e65.

[132] Law CK. Combustion physics. New York: Cambridge University Press; 2006.[133] Shaddix CR, Molina A. Particle imaging of ignition and devolatilization of

pulverized coal during oxy-fuel combustion. Proceedings of the CombustionInstitute 2009;32:2091e8.

[134] Fan Y, Zou Z, Cao Z, Xu Y, Jiang X. Ignition characteristics of pulverized coalunder high oxygen concentrations. Energy & Fuels 2008;22:892e7.

[135] Molina A, Shaddix CR. Ignition and devolatilization of pulverized bituminouscoal particles during oxygen/carbon dioxide coal combustion. Proceedings ofthe Combustion Institute 2007;31:1905e12.

[136] Stivers C, Levendis YA. Ignition of sigle coal particles in O2/N2/CO2 atmo-spheres. In: The 35th international technical conference on clean coal & fuelsystems. Clearwater, FL; 2010.

[137] Taniguchi M, Kobayashi H, Kiyama K, Shimogori Y. Comparison of flamepropagation properties of petroleum coke and coals of different rank. Fuel2009;88:1478e84.

[138] Kee RJ, Rupley FM, Miller JA, Coltrin ME, Grcar JF, Meeks E, et al. CHEMKINrelease 4.0. San Diego, CA: Reaction Design, Inc.; 2004.

[139] Smith GP, Golden DM, Frenklach M, Moriarty NW, Eiteneer B, Goldenberg M,et al. GRI-Mech 3.0, http://www.me.berkeley.edu/gri_mech/; 2009.

[140] Molina A, Hecht ES, Shaddix CR. Ignition of a group of coal particles inoxyfuel combustion with CO2 recirculation. In: The 34th internationaltechnical conference on coal utilization and fuel systems. Clearwater, FL,USA; 2009.

[141] Levendis YA. Personal communication; 2011.[142] Qiao Y, Zhang L, Binner E, Xu M, Li C- Z. An investigation of the causes of the

difference in coal particle ignition temperature between combustion in airand in O2/CO2. Fuel 2010;89:3381e7.

[143] Bejarano PA, Levendis YA. Single-coal-particle combustion in O2/N2 and O2/CO2 environments. Combustion and Flame 2008;153:270e87.

[144] Harris DJ, Smith IW. Intrinsic reactivity of petroleum coke and brown coalchar to carbon dioxide, steam and oxygen. Symposium (International) onCombustion 1991;23:1185e90.

[145] Mann AP, Kent JH. A computational study of heterogeneous char reactions ina full-scale furnace. Combustion and Flame 1994;99:147e56.

[146] Stanmore BR, Visona SP. The contribution to char burnout from gasificationby H2O and CO2 during pulverized-coal flame combustion. Combustion andFlame 1998;113:274e6.

[147] Kajitani S, Hara S, Matsuda H. Gasification rate analysis of coal char witha pressurized drop tube furnace. Fuel 2002;81:539e46.

[148] Shaddix CR, Molina A. Effect of O2 and high CO2 concentrations on PC charburning rates during oxy-fuel combustion. In: The 33rd international tech-nical conference on coal utilization and fuel systems. Clearwater, FL, USA;2008.

[149] Shaddix CR, Hecht ES, Geier M, Molina A, Haynes BS. Effect of gasificaitonreactions on oxy-fuel combustion of pulverized coal char. In: The 35thinternational technical conference on clean coal & fuel systems. Clearwater,FL; 2010.

[150] Varhegyi G, Szabo P, Jakab E, Till F, Richard JR. Mathematical modeling of charreactivity in Ar-O2 and CO2-O2 mixtures. Energy Fuels 1996;10:1208e14.

[151] Varhegyi G, Till F. Comparison of temperature-Programmed char combustionin CO2-O2 and Ar-O2 mixtures at elevated pressure. Energy Fuels 1999;13:539e40.

[152] Liu H. Combustion of coal chars in O2/CO2 and O2/N2 mixtures: a compara-tive study with non-isothermal thermogravimetric analyzer (TGA) tests.Energy & Fuels 2009;23:4278e85.

[153] Liu H. A comparison of combustion of coal chars in O2/CO2 and O2/N2mixtures - Isothermal TGA studies. International Journal of Reactor Engi-neering 2009;7.

[154] Annamalai K, Puri IK. Combustion science and engineering. Boca Raton: CRCPress/Taylor & Francis; 2007.

[155] Mitchell RE, Kee RJ, Glarborg P, Coltrin ME. The effect of CO conversion in theboundary layers surrounding pulverized-coal char particles. Symposium(International) on Combustion 1991;23:1169e76.

[156] Murphy JJ, Shaddix CR. Combustion kinetics of coal chars in oxygen-enrichedenvironments. Combustion and Flame 2006;144:710e29.

[157] Tognotti L, Longwell JP, Sarofim AF. The products of the high temperatureoxidation of a single char particle in an electrodynamic balance. Symposium(International) on Combustion 1991;23:1207e13.

[158] Field MA. Rate of combustion of size-graded fractions of char from a low-rank coal between 1 200 K and 2 000 K. Combustion and Flame 1969;13:237e52.

[159] Hecht ES, Shaddix CR, Molina A, Haynes BS. Effect of CO2 gasification reactionon oxy-combustion of pulverized coal char. Proceedings of the CombustionInstitute 2011;33:1699e706.

[160] Guo X, Tay HL, Zhang S, Li C- Z. Changes in char structure during the gasi-fication of a Victorian brown coal in steam and oxygen at 800 oC. Energy &Fuels 2008;22:4034e8.

[161] Smith IW. The combustion rates of coal chars: a review. Symposium(International) on Combustion 1982;19:1045e65.

[162] Toporov D, Bocian P, Heil P, Kellermann A, Stadler H, Tschunko S, et al.Detailed investigation of a pulverized fuel swirl flame in CO2/O2 atmosphere.Combustion and Flame 2008;155:605e18.

[163] Brown BW, Smoot LD, Smith PJ, Hedman PO. Measurement and prediction ofentrained-flow gasification processes. AIChE Journal 1988;34:435e46.

[164] Incropera FP, DeWitt DP. Fundamentals of heat and mass transfer. 5th ed.New York: Wiley; 2002.

[165] Niksa S, Liu G-s, Hurt RH. Coal conversion submodels for design applicationsat elevated pressures. Part I. devolatilization and char oxidation. Progress inEnergy and Combustion Science 2003;29:425e77.

[166] Liu G-S, Niksa S. Coal conversion submodels for design applications atelevated pressures. Part II. Char gasification. Progress in Energy andCombustion Science 2004;30:679e717.

[167] Scherello A, Hackert G, Wirtz S, Bonn B, Seewald H. Drop tube investigationsinto coal particle combustion at elevated temperatures and pressures. In:Ninth international conference on coal science. DGMK Tagungsberichte97031997: 99e1002.

[168] Suriyawong A, Gamble M, Lee M-H, Axelbaum R, Biswas P. Submicrometerparticle formation and mercury speciation under O2-CO2 coal combustion.Energy & Fuels 2006;20:2357e63.

[169] Stam A, Ploumen P, Brem G. Ash related aspects of oxy-combustion of coaland biomass: a thermodynamic approach. In: The 34th international tech-nical conference on coal utilization and fuel systems. Clearwater, Florida,USA; 2009.

[170] Sheng C, Li Y. Experimental study of ash formation during pulverized coalcombustion in O2/CO2 mixtures. Fuel 2008;87:1297e305.

[171] Sheng C, Lu Y, Gao X, Yao H. Fine ash formation during pulverized coalcombustion A comparison of O2/CO2 combustion versus air combustion.Energy & Fuels 2007;21:435e40.

[172] Lighty JS, Veranth JM, Sarofim AF. Combustion aerosols: factors governingtheir size and composition and implications to human healthy. Journal of theAir & Waste Management Association 2000;50:1565e618.

[173] Richards GH, Slater PN, Harb JN. Simulation of ash deposit growth ina pulverized coal-fired pilot scale reactor. Energy & Fuels 1993;7:774e81.

[174] Li S, Wu Y, Whitty KJ. Ash deposition behavior during chareslag transitionunder simulated gasification conditions. Energy & Fuels 2010;24:1868e76.

[175] Fryda L, Sobrino C, Cieplik M, van de Kamp WL. Study on ash depositionunder oxyfuel combustion of coal/biomass blends. Fuel 2010;89:1889e902.

[176] Yong SZ, Gazzino M, Ghoniem A. Modeling the slag layer in solid fuel gasi-fication and combustion e formulation and sensitivity analysis. Fuel 2012;92:162e70.

[177] Strandstrom K, Mueller C, Hupa M. Development of an ash particle deposi-tion model considering build-up and removal mechanisms. Fuel ProcessingTechnology 2007;88:1053e60.

[178] Senda J, Kobayashi M, Iwashita S, Fujimoto H. Modeling of diesel sprayimpinging on flat wall. JSME International Journal 1996;39(4).

[179] Normann F, Andersson K, Leckner B, Johnsson F. Emission control of nitrogenoxides in the oxy-fuel process. Progress in Energy and Combustion Science2009;35:385e97.

[180] Hu Y, Naito S, Kobayashi N, Hasatani M. CO2, NOx and SO2 emissions fromthe combustion of coal with high oxygen concentration gases. Fuel 2000;79:1925e32.

[181] Okazaki K, Ando T. NOx reduction mechanism in coal combustion withrecycled CO2. Energy 1997;22:207e15.

[182] Liu H, Zailani R, Gibbs BM. Comparisons of pulverized coal combustion in airand in mixtures of O2/CO2. Fuel 2005;84:833e40.

[183] Liu H, Zailani R, Gibbs BM. Pulverized coal combustion in air and in O2/CO2mixtures with NOx recycle. Fuel 2005;84:2109e15.

Page 58: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214 213

[184] Hu YQ, Kobayashi N, Hasatani M. The reduction of recycled-NOx in coalcombustion with O2/recycled flue gas under low recycling ratio. Fuel 2001;80:1851e5.

[185] Kiga T, Takano S, Kimura N, Omata K, Okawa M, Mori T, et al. Characteristicsof pulverized-coal combustion in the system of oxygen/recycled flue gascombustion. Energy Conversion and Management 1997;38:S129e34.

[186] Stanger R, Wall T. Sulphur impacts during pulverised coal combustion inoxy-fuel technology for carbon capture and storage. Progress in Energy andCombustion Science 2011;37:69e88.

[187] Zheng L, Furimsky E. Assessment of coal combustion in O2þCO2 by equi-librium calculations. Fuel Processing Technology 2003;81:23e34.

[188] Glarborg P, Bentzen LLB. Chemical effects of a high CO2 concentration in oxy-fuel combustion of methane. Energy & Fuels 2008;22:291e6.

[189] Rehfeldt S, Kuhr C, Ehmann M, Bergins C, Scheffknecht G, Maier J, et al. Basicexperiments and CFD calculations of air and oxyfuel firing of lignite andbituminous coals in 0.5 and 1 MW scale combustion test facilities. In: The34th international technical conference on clean coal and fuel systems.Clearwater, FL; 2009.

[190] Andersson K, Normann F, Johnsson F, Leckner B. NO emission during oxy-fuel combustion of lignite. Industrial & Engineering Chemistry Research2008;47:1835e45.

[191] Kimura N, Omata K, Kiga T, Takano S, Shikisima S. The characteristics ofpulverized coal combustion in O2/CO2 mixtures for CO2 recovery. EnergyConversion and Management 1995;36:805e8.

[192] Muñiz L, Mungal MG. Instantaneous flame-stabilization velocities in lifted-jet diffusion flames. Combustion and Flame 1997;111:16e30.

[193] Williams FA. Combustion theory: the fundamental theory of chemicallyreacting flow systems. 2nd ed. Menlo Park, Calif.: Benjamin/Cummings Pub.Co.; 1985.

[194] Kalghatgi GT. Lift-off heights and visible lengths of vertical turbulent jet diffu-sion flames in still air. Combustion Science and Technology 1984;41:17e29.

[195] Kalghatgi GT. Blow-out stability of gaseous jet diffusion flames. Part I: in stillair. Combustion Science and Technology 1981;26:233e9.

[196] Spalding DB. Combustion and mass transfer: a textbook with mulitple-choice exercises for engineering students. 1st ed. Oxford/New York: Perga-mon Press; 1979.

[197] Suda T, Masuko K, Ji Sato, Yamamoto A, Okazaki K. Effect of carbon dioxideon flame propagation of pulverized coal clouds in CO2/O2 combustion. Fuel2007;86:2008e15.

[198] Liu F, Guo H, Smallwood GJ. The chemical effect of CO2 replacement of N2 inair on the burning velocity of CH4 and H2 premixed flames. Combustion andFlame 2003;133:495e7.

[199] Vagelopoulos CM, Egolfopoulos FN, Law CK. Further considerations on thedetermination of laminar flame speeds with the counterflow twin-flametechnique. Symposium (International) on Combustion 1994;25:1341e7.

[200] Vagelopoulos CM, Egolfopoulos FN, Law CK. Laminar flame speeds ofmethane at 1 atm, phi ¼ 0.4e1.7, with three bath gases: Ar, N2, and CO2.http://www.me.berkeley.edu/gri_mech/version30/text30.html.

[201] Chui EH, Douglas MA, Tan Y. Modeling of oxy-fuel combustion for a westernCanadian sub-bituminous coal. Fuel 2003;82:1201e10.

[202] Hohenwarter U, Griendl L. Design of a novel oxycoal burner for controlledstaged combustion. In: 32nd international technical conference on coalutilization & fuel systems. Clearwater, FL; 2007.

[203] Griendl L, Hohenwarter U, Karl J. CFD aided Design of an Oxycoal Test Burnerwith High Recirculation Rates. In: 33rd International Technical Conferenceon Coal Utilization & Fuel Systems. Clearwater, Florida; 2008.

[204] Bool L, Rosen L, Laux S, Kobayashi H. Direct oxygen injection in pulverized coalburners for flame stability and pollutant control. In: The 35th InternationalTechnical Conference on Clean Coal & Fuel Systems. Clearwater, FL; 2010.

[205] Zhang J, Kelly KE, Eddings EG, Wendt JOL. Ignition in 40 kW co-axialturbulent diffusion oxy-coal jet flames. Proceedings of the CombustionInstitute 2011;33:3375e82.

[206] Fry A, Adams B, Shan J. Oxy-burner retrofit principles for existing coal-firedutility boilers. In: The 35th international technical conference on clean coal &fuel systems. Clearwaer, FL; 2010.

[207] Khare SP, Wall TF, Farida AZ, Liu Y, Moghtaderi B, Gupta RP. Factors influ-encing the ignition of flames from air-fired swirl pf burners retrofitted tooxy-fuel. Fuel 2008;87:1042e9.

[208] Tigges KD, Klauke F, Bergins C, Busekrus K, Niesbach J, Ehmann M, et al.Oxyfuel combustion retrofits for existing power stations. In: The 33rdinternational technical conference on coal utilization & fuel systems. Clear-water, Florida, USA; 2008.

[209] Tigges KD, Klauke F, Bergins C, Busekrus K, Niesbach J, Ehmann M, et al.Conversion of existing coal-fired power plants to oxyfuel combustion: casestudy with experimental results and CFD-simulations. Energy Procedia 2009;1:549e56.

[210] Heil P, Toporov D, Stadler H, Tschunko S, Forster M, Kneer R. Development ofan oxycoal swirl burner operating at low O2 concentrations. Fuel 2009;88:1269e74.

[211] Suda T, Takafuji M, Hirata T, Yoshino M, Sato J. A study of combustionbehavior of pulverized coal in high-temperature air. Proceedings of theCombustion Institute 2002;29:503e9.

[212] Saponaro A, Cumbo D, Gazzino M. Mild coal-combustion technology forFuture zero emission power plant. In: Italian section of the combustioninstitute 30th meeting on combustion. Ischia, Italy; 2007.

[213] Wendt JF, Anderson JD, Von Karman Institute for Fluid Dynamics. Compu-tational fluid dynamics: an introduction. 2nd ed. Berlin/New York: Springer;1996.

[214] Migdal D, Agosta VD, American Society of Mechanical Engineers. A sourceflow model for continuum gas-particle flow. New York, N.Y.: ASME; 1967.

[215] Muller M, Lemp O, Leiser S, Schnell U, Grathwohl S, Maier J, et al. Advancedmodeling of pulverized coal combustion under oxy-fuel conditions. In: The35th International Technical Conference on Clean Coal and Fuel Systems.Clearwater, FL; 2010.

[216] FLUENT 6.3 user’s guide. Fluent Inc.; 2006.[217] Edge P, Gubba SR, Ma L, Porter R, Pourkashanian M, Williams A. LES

modelling of air and oxy-fuel pulverised coal combustion - impact on flameproperties. Proceedings of the Combustion Institute 2011;33:2709e16.

[218] Porter R, Liu F, Pourkashanian M, Williams A, Smith D. Evaluation of solutionmethods for radiative heat transfer in gaseous oxy-fuel combustion envi-ronments. Journal of Quantitative Spectroscopy and Radiative Transfer 2010;111:2084e94.

[219] SmithTF, ShenZF, Friedman JN. Evaluationof coefficients for theweighted sumof gray gases model. The ASME Journal of Heat Transfer 1982;104:602e8.

[220] Viskanta R, Menguc MP. Radiation heat transfer in combustion systems.Progress in Energy and Combustion Science 1987;13:97e160.

[221] Edwards DK, MenardWA. Comparison of models for correlation of total bandabsorption. Applied Optics 1964;3:621e5.

[222] Johansson R, Andersson K, Leckner B, Thunman H. Models for gaseousradiative heat transfer applied to oxy-fuel conditions in boilers. InternationalJournal of Heat and Mass Transfer 2010;53:220e30.

[223] Rehfeldt S, Kuhr C, Ehmann M, Bergins C, Tigges KD, Wu S. Modeling ofradiative heat transfer in an oxyfuel atmosphere with a wide band modeland a weighted sum of gray gases. In: The 35th international technicalconference on clean coal and fuel systems. Clearwater, FL; 2010.

[224] Nikolopoulos N, Nikolopoulos A, Karampinis E, Grammelis P, Kakaras E.Numerical investigation of the oxy-fuel combustion in large scale boilersadopting the ECO-Scrub technology. Fuel 2011;90:198e214.

[225] Arthur JR. Reactions between carbon and oxygen. Transactions of theFaraday Society 1951;47:164e78.

[226] Kuhr C, Ehmann M, Rehfeldt S, Bergins C, Maier J, Scheffknecht G, et al.Modeling of char combustion in CO2/O2 and N2/O2 atmospheres. In: The 35thinternational technical conference on clean coal and fuel systems. Clear-water, FL; 2010.

[227] Vascellari M, Cau G. Numerical simulation of pulverized coal oxy-combustion with exhaust gas recirculation. In: Proceeding of CCT2009Fourth International Conference on Clean Coal Technologies. Dresden, Ger-many; 2009.

[228] Zhang M, Yu J, Xu X. A new flame sheet model to reflect the influence of theoxidation of CO on the combustion of a carbon particle. Combustion andFlame 2005;143:150e8.

[229] Westbrook CK, Dryer FL. Simplified reaction mechanisms for the oxidation ofhydrocarbon fuels in flames. Combustion Science and Technology 1981;27:31e43.

[230] Jones WP, Lindstedt RP. Global reaction schemes for hydrocarbon combus-tion. Combustion and Flame 1988;73:233e49.

[231] Yin C, Kar SK, Rosendahl L, Hvid SL. Co-firing straw with coal in a swirl-stabilized dual-feed burner: modelling and experimental validation. Bio-resource Technology 2010;101:4169e78.

[232] Kumar M, Zhang C, Monaghan RFD, Singer SL, Ghoniem AF. CFD simulation ofentrained flow gasification with improved devolatilization and charconsumption submodels. In: ASME 2009 international mechanical engi-neering congress & exposition. Lake Buena Vista, FL; 2009.

[233] Andersen J, Rasmussen CL, Giselsson T, Glarborg P. Global combustionmechanisms for use in CFD modeling under oxy-fuel conditions. Energy &Fuels 2009;23:1379e89.

[234] Magnussen BF, Hjertager BH. On mathematical modeling of turbulentcombustion with special emphasis on soot formation and combustion.Symposium (International) on Combustion 1977;16:719e29.

[235] Spalding DB. Mixing and chemical reaction in steady confined turbulentflames. Symposium (International) on Combustion 1971;13:649e57.

[236] Breussin F, Lallemant N, Weber R. Computing of oxy-natural gas flames usingboth a global combustion scheme and a chemical equilibrium procedure.Combustion Science and Technology 2000;160:369e97.

[237] Magnussen BF. On the structure of turbulence and a generalized Eddydissipation concept for chemical reaction in turbulent flow. In: 19th AIAAAerospace Science Meeting. St.Louis, Missouri; 1981.

[238] Brink A, Kilpinen P, Hupa M, Kjäldman L. Study of alternative descriptions ofmethane oxidation for CFD modeling of turbulent combustors. J CombustionScience and Technology 1999;141:59e81.

[239] Kim JP, Schnell U, Scheffknecht G. Comparison of different global reactionmechanisms for MILD combustion of natural gas. Combustion Science andTechnology 2008;180:565e92.

[240] Fortsch D, Kluger F, Schnell U, Spliethoff H, Hein KRG. A kinetic model for theprediction of no emissions from staged combustion of pulverized coal.Symposium (International) on Combustion 1998;27:3037e44.

[241] Kluger F, Moenckert P, Stamatelopoulos G-N. Alstom’s oxy-combustiontechnology development - update on pilot plants operation. In: 35th inter-national technical conference on clean coal & fuel systems. Clearwater, FL;2010.

Page 59: Progress in Energy and Combustion Science · 2019-12-31 · Post-combustion capture:CO2 is separated from the flue gases usingchemicalsolvents[7],sorbents(suchascalciumoxide[8] orcarbonfibers[9])andmembranes[10]withoutchangingthe

L. Chen et al. / Progress in Energy and Combustion Science 38 (2012) 156e214214

[242] Sturgeon DW, Cameron ED, Fitzgerald FD. Demonstration of an oxyfuelcombustion system. Energy Procedia 2009;1:471e8.

[243] Strömberg L, Lindgren G, Jacoby J, Giering R, Anheden M, Burchhardt U, et al.Update on Vattenfall’s 30 MWth oxyfuel pilot plant in Schwarze Pumpe.Energy Procedia 2009;1:581e9.

[244] Cameron ED, Sturgeon DW, McGhie C, Fitzgerald FD. Demonstration of thedoosan babcock 40 MWt oxycoal combustion system. In: The 35th Interna-tional Technical Conference onCleanCoal&Fuel SystemsClearwater, FL; 2010.

[245] Chun B-S, Wilkinson GT. Interfacial tension in high-pressure carbon dioxidemixtures. Industrial & Engineering Chemistry Research 1995;34:4371e7.

[246] IAPWS release on surface tension of ordinary water SUBSTANCE. Interna-tional Association for the Properties of Water and Steam; 1994.

[247] Zhu DL, Egolfopoulos FN, Law CK. Experimental and numerical determina-tion of laminar flame speeds of methane/(Ar, N2, CO2)-air mixtures asfunction of stoichiometry, pressure, and flame temperature. Symposium(International) on Combustion 1989;22:1537e45.

[248] Westbrook CK, Dryer FL. Chemical kinetic modeling of hydrocarboncombustion. Progress in Energy and Combustion Science 1984;10:1e57.

[249] Chui EH, Majeski AJ, Douglas MA, Tan Y, Thambimuthu KV. Numericalinvestigation of oxy-coal combustion to evaluate burner and combustordesign concepts. Energy 2004;29:1285e96.

[250] Weller AE, Rising BW, Boiarski AA, Nordstrom RJ, Barrett RE, Luce RG.Experimental evaluation of firing pulverized coal in a CO2/O2 atmosphere.Argonne National Laboratory; 1985. Report ANL/CNSV-TM-168342.

[251] Law CK, Williams FA. Kinetics and convection in the combustion of alkanedroplets. Combustion and Flame 1972;19:393e405.

[252] Smith PJ, Smoot LD. One-dimensional model for pulverized coal combustionand gasification. Combustion Science and Technology 1980;23:17e31.

[253] Patankar SV, Spalding DB. A computer model for three-dimensional flow infurnaces. Symposium (International) on Combustion 1973;14:605e14.

[254] Lockwood FC, Salooja AP, Syed SA. A prediction method for coal-firedfurnaces. Combustion and Flame 1980;38:1e15.

[255] Dryer FL, Glassman I. High-temperature oxidation of CO and CH4. Sympo-sium (International) on Combustion 1973;14:987e1003.

[256] Leckner B. Spectral and total emissivity of water vapor and carbon dioxide.Combustion and Flame 1972;19:33e48.