+ All Categories
Home > Documents > CFD simulation of entrained-flow coal gasification: Coal particle density/sizefraction effects

CFD simulation of entrained-flow coal gasification: Coal particle density/sizefraction effects

Date post: 18-Nov-2023
Category:
Upload: independent
View: 0 times
Download: 0 times
Share this document with a friend
11
CFD simulation of entrained-ow coal gasication: Coal particle density/size fraction effects Andrew Slezak a,b , John M. Kuhlman a,b, , Lawrence J. Shadle a , James Spenik a,c , Shaoping Shi a,d a National Energy Technology Laboratory, Morgantown, WV 26507-0880, United States b West Virginia University, Mechanical and Aerospace Eng. Dept. Morgantown, WV 26506-6106, United States c REM Engineering Services, Morgantown, WV 26505, United States d Ansys, Inc., Morgantown, WV 26505, United States abstract article info Article history: Received 21 August 2009 Accepted 18 February 2010 Available online 8 April 2010 Keywords: Coal gasication Entrained-ow reactor CFD Discrete Phase Method DPM Particlewall interactions Coal size/density fractions Computational Fluid Dynamics (CFD) simulation of commercial-scale two-stage upow and single-stage downow entrained-ow gasiers was conducted to study effects of simulating both the coal particle density and size variations. A previously-developed gasication CFD model was modied to account for coal particle density and size distributions as produced from a typical rod mill. Postprocessing tools were developed for analysis of particlewall impact properties. For the two-stage upow gasier, three different simulations are presented: two (Case 1 and Case 2) used the same devolatilization and char conversion models from the literature, while Case 3 used a different devolatilization model. The Case 1 and Case 3 solutions used average properties of a Pittsburgh #8 seam coal (d = 108 μm, SG = 1.373), while Case 2 was obtained by injecting and tracking all of the series of 28 different coal particle density and size mass fractions obtained by colleagues at PSU as a part of the current work, for this same coal. Simulations using the two devolatilization models (Case 1 and Case 3) were generally in reasonable agreement. Differences were observed between the single-density solution and the density/size partitioned solution (Case 1 and Case 2). The density/size partitioned solution predicted nominally 10% less CO and over 5% more H 2 by volume in the product gas stream. Particle residence times and trajectories differed between these two solutions for the larger density/size fractions. Fixed carbon conversion was 4.3% higher for the partitioned solution. Particlewall impact velocities did not vary greatly. Grid independence studies for the two-stage upow gasier geometry showed that the grid used in the comparison studies was adequate for predicting exit gas composition and wall impact velocities. Validation studies using experimental data for the Pittsburgh #8 coal from the SRI International pressurized coal ow reactor (PCFR) at 30 atmospheres indicated adequate agreement for gasication and combustion cases, but poor agreement for a pyrolysis case. Simulation of a single-stage downow gasier yielded an exit gas composition that was in reasonable agreement with published data. © 2010 Elsevier B.V. All rights reserved. 1. Introduction The present study has been undertaken to demonstrate capability to predict the performance of entrained-ow gasiers using current state-of-the-art Computational Fluid Dynamics (CFD) software, and more specically to predict differences within the gasier in behaviors (e.g., particle residence times, char conversion, and trajectories) of various size and density fractions of the pulverized coal fuel produced from a typical rod mill. This effort is a part of a larger study, the Coal Partitioning Project, or CPP, which is being supported by the DOE National Energy Technology Laboratory (NETL). In the CPP project, a Pittsburgh #8 seam coal from the Bailey mine in southwestern Pennsylvania has been ground using a METSO rod mill to specica- tions typical of slurry-fed entrained-ow gasiers and separated into four density fractions, and then each of these density fractions has been further classied into seven diameter ranges [1,2]. The parent coal and each of the 28 resulting density/size cuts have been characterized by an ultimate and proximate analysis, a maceral analysis, and a mineral analysis of the ash from each cut [1,2]. While a density separation is known to separate hydrogen-rich, liptinite macerals in the lighter fractions from denser wood-derived vitrinite, and denser still mineral impregnated inertinite macerals, there was little evidence that this was the case in this vitrinite-rich coal. The density separation however did produce a separation in the mineral matter such that the lighter fractions consisted of ne clays imbedded within the coal particles, the denser fractions had subsequently higher concentrations of clays and heavier pyrite minerals, while the densest fraction was essentially individual minerals with some organic Powder Technology 203 (2010) 98108 Corresponding author. National Energy Technology Laboratory, Morgantown, WV 26507-0880, United States. Tel.: +1 304 293 3180; fax: +1 304 293 6689. E-mail address: [email protected] (J.M. Kuhlman). 0032-5910/$ see front matter © 2010 Elsevier B.V. All rights reserved. doi:10.1016/j.powtec.2010.03.029 Contents lists available at ScienceDirect Powder Technology journal homepage: www.elsevier.com/locate/powtec
Transcript

Powder Technology 203 (2010) 98–108

Contents lists available at ScienceDirect

Powder Technology

j ourna l homepage: www.e lsev ie r.com/ locate /powtec

CFD simulation of entrained-flow coal gasification: Coal particle density/sizefraction effects

Andrew Slezak a,b, John M. Kuhlman a,b,⁎, Lawrence J. Shadle a, James Spenik a,c, Shaoping Shi a,d

a National Energy Technology Laboratory, Morgantown, WV 26507-0880, United Statesb West Virginia University, Mechanical and Aerospace Eng. Dept. Morgantown, WV 26506-6106, United Statesc REM Engineering Services, Morgantown, WV 26505, United Statesd Ansys, Inc., Morgantown, WV 26505, United States

⁎ Corresponding author. National Energy Technology26507-0880, United States. Tel.: +1 304 293 3180; fax:

E-mail address: [email protected] (J.M. K

0032-5910/$ – see front matter © 2010 Elsevier B.V. Adoi:10.1016/j.powtec.2010.03.029

a b s t r a c t

a r t i c l e i n f o

Article history:Received 21 August 2009Accepted 18 February 2010Available online 8 April 2010

Keywords:Coal gasificationEntrained-flow reactorCFDDiscrete Phase MethodDPMParticle–wall interactionsCoal size/density fractions

Computational Fluid Dynamics (CFD) simulation of commercial-scale two-stage upflow and single-stagedownflow entrained-flow gasifiers was conducted to study effects of simulating both the coal particledensity and size variations. A previously-developed gasification CFD model was modified to account for coalparticle density and size distributions as produced from a typical rod mill. Postprocessing tools weredeveloped for analysis of particle–wall impact properties.For the two-stage upflow gasifier, three different simulations are presented: two (Case 1 and Case 2) usedthe same devolatilization and char conversion models from the literature, while Case 3 used a differentdevolatilization model. The Case 1 and Case 3 solutions used average properties of a Pittsburgh #8 seam coal(d=108 μm, SG=1.373), while Case 2 was obtained by injecting and tracking all of the series of 28 differentcoal particle density and size mass fractions obtained by colleagues at PSU as a part of the current work, forthis same coal. Simulations using the two devolatilization models (Case 1 and Case 3) were generally inreasonable agreement. Differences were observed between the single-density solution and the density/sizepartitioned solution (Case 1 and Case 2). The density/size partitioned solution predicted nominally 10% lessCO and over 5% more H2 by volume in the product gas stream. Particle residence times and trajectoriesdiffered between these two solutions for the larger density/size fractions. Fixed carbon conversion was 4.3%higher for the partitioned solution. Particle–wall impact velocities did not vary greatly.Grid independence studies for the two-stage upflow gasifier geometry showed that the grid used in thecomparison studies was adequate for predicting exit gas composition and wall impact velocities. Validationstudies using experimental data for the Pittsburgh #8 coal from the SRI International pressurized coal flowreactor (PCFR) at 30 atmospheres indicated adequate agreement for gasification and combustion cases, butpoor agreement for a pyrolysis case. Simulation of a single-stage downflow gasifier yielded an exit gascomposition that was in reasonable agreement with published data.

Laboratory, Morgantown, WV+1 304 293 6689.uhlman).

ll rights reserved.

© 2010 Elsevier B.V. All rights reserved.

1. Introduction

The present study has been undertaken to demonstrate capabilityto predict the performance of entrained-flow gasifiers using currentstate-of-the-art Computational Fluid Dynamics (CFD) software, andmore specifically to predict differenceswithin the gasifier in behaviors(e.g., particle residence times, char conversion, and trajectories) ofvarious size and density fractions of the pulverized coal fuel producedfrom a typical rod mill. This effort is a part of a larger study, the CoalPartitioning Project, or CPP, which is being supported by the DOENational Energy Technology Laboratory (NETL). In the CPP project, aPittsburgh #8 seam coal from the Bailey mine in southwestern

Pennsylvania has been ground using a METSO rod mill to specifica-tions typical of slurry-fed entrained-flow gasifiers and separated intofour density fractions, and then each of these density fractions hasbeen further classified into seven diameter ranges [1,2]. The parentcoal and each of the 28 resulting density/size cuts have beencharacterized by an ultimate and proximate analysis, a maceralanalysis, and a mineral analysis of the ash from each cut [1,2]. While adensity separation is known to separate hydrogen-rich, liptinitemacerals in the lighter fractions from denser wood-derived vitrinite,and denser still mineral impregnated inertinite macerals, there waslittle evidence that this was the case in this vitrinite-rich coal. Thedensity separation however did produce a separation in the mineralmatter such that the lighter fractions consisted of fine clays imbeddedwithin the coal particles, the denser fractions had subsequently higherconcentrations of clays and heavier pyrite minerals, while the densestfraction was essentially individual minerals with some organic

Table 2Proximate and ultimate analyses of Pittsburgh #8 coal from the Bailey mine [1,2].

Component Pittsburgh #8 coal

Fixed carbon 57.69Volatiles 33.52Ash 7.79Moisture 1C 76.83H 5.49O 3.91N 1.4S 1.46Cl 0

Fig. 1. Commercial-scale, two-stage, upflow, slurry-fed, entrained-flow gasifiergeometry and grid.

Table 1Chronological summary of previous coal gasification CFD studies.

Ref # Authors Year Code used Methodused

Configurations(s)simulated

[12] Boumaet al.

1999 CFX 4.2 Eulerian–Eulerian

Axisymmetric entrained-flowgasification simulator

[13] Chenet al.

1999 – DPM 200 T/day 2-stage HYCOLgasifier

[14] Shanamet al.

2000 Fluent 5.4 DPM N. Dakota EERCtransport gasifier

[15] Fletcheret al.

2000 CFX 4 DPM 1 MW entrained-flowbiomass gasifier

[16] Chenet al.

2000 – DPM 200 T/day 2-stageHYCOL gasifier

[7] Chenet al.

2000 – DPM 200 T/day 2-stageHYCOL gasifier

[17] Tominagaet al.

2000 Fluent DPM 2 T/day and 50 T/day2-stage HYCOL gasifiers

[8] Chenet al.

2001 - DPM 200 T/day 2-stageHYCOL gasifier

[9] Liu et al. 2001 ModifiedΨ–ω

DPM Bench-scale KIERentrained-flow gasifier

[18] Bockelieet al.

2002 GLACIER DPM 2-Stage upflow and1-stage downflow gasifiers

[19] Liu andKojima

2004 – DPM Pilot-scale 2-stageHYCOL gasifier

[20] Liu andKojima

2004 – DPM pIlot-scale 2-stageHYCOL gasifier

[21] Rouciset al.

2004 CD-adapco

– Pilot-scale gasifier

[22] Watanabeet al.

2004 CFX 4 DPM 2 T/day entrained-flow gasifier

[23] Guentherand Breault

2005 MFIX Eulerian–Eulerian

Transport gasifier

[3] Shi et al. 2006 Fluent 6.1 DPM Commercial 2-stage, upflowentrained-flow gasifier

[24] Tominagaet al.

2006 Fluent andRESORT

DPM 2 T/day and 50 T/day2-stage HYCOL gasifiers

[25] Lu et al. 2008 – DPM Lime-injected fluidized-bedcoal gasifier

[26] Forstneret al.

2006 Fluent 6.1 DPM Biomass-fueled, fixed-bedfurnace

[27] Syred et al. 2007 Fluent 6.1 DPM 500 kV coal-fired furnace

Table 3Composite Bailey coal specific gravity (SG) and particle size (PS) [1,2].

Gravity fraction BSG1PS0 B

Avg. specific gravity 1.2a

wt.% 47.84 4

Size class PS1 PS2 PS3

Avg. diameter, μm 800a 512 318BSG1 wt.% 5.31 10.85 31.15BSG2 wt.% 3.91 9.53 21.66BSG3 wt.% 16.15 11.5 23.38BSG4 wt.% 8.69 5.93 17.5

Composite SG=1.373; composite Sauter-mean diameter=108 μm.a Assumed value.

99A. Slezak et al. / Powder Technology 203 (2010) 98–108

impurities [2]. The CPP project objectives are to evaluate the influenceof these compositional differences in coal particles on the partitioningof inorganics produced during gasification into flyash and slag. Assuch it is important to understand the differences in the trajectories ofthese various particles and their relative tendencies to strike the wall.

The main focus of the present CFD simulations is a commercial-scale, two-stage, upflow, slurry-fed, entrained-flow gasifier geometrypreviously studied by Shi et al. [3]. The Fluent user-defined function(UDF) developed by Shi to simulate gasification has been extendedherein to allow the feeding of (28) multiple coal fractions into thegasifier, each at the appropriate mass flow rate and chemicalcomposition. Results have been obtained at a typical operatingcondition using the devolatilization chemistry model of Kobayashiet al. [4] and the shrinking core char conversion models of Wen andChaung [5]. Results are compared for one solution for the parent coalat themeasured Sauter-mean diameter of 108 μmand average densityof 1373 kg/m3 for the parent coal, and for a second solution for theinjection of all 28 density/size cuts, each at the appropriate injectionmass flow rate, and each with its own volatile chemical compositionand ultimate yield as governed by the measured ultimate andproximate analysis for that size and density fraction, as given by theKobayashi et al. devolatilization model. Also, a solution using thedevolatilization model from PC Coal Lab6 has been compared with thefirst solution.

2. Background

Some previous CFD gasifier simulations have investigated thebehavior of a range of coal particle size fractions (e.g., [7–9]), but noneto date have included various density fractions. This is at least partly

SG2PS0 BSG3PS0 BSG4PS0

1.45 2.1 3.3a

7.57 3.46 1.12

PS4 PS5 PS6 PS7

181 128 90 50a

13.36 11.57 10.79 16.979.15 7.38 6.99 41.39

10.71 9.12 8.79 20.48.31 8.65 13.49 37.43

Table 4Assumed model conditions for present solutions.

Case 1 Case 2 Case 3

Devolatilization Kobayashi Kobayashi PC Coal Laba

Char conversion Wen/Chaung Wen/Chaung Wen/ChaungCoal particle(s) d=108 μm 28 CPP size/ρ cuts d=108 μmCp equations Polynomial Polynomial Polynomialk equations Power law Power law Power law

a Recommended one point tuning of this model using experimental entrained-flowdata has not yet been completed.

Table 5Comparison of average temperatures, FC conversion, and exit gas compositions (byvol.).

Case 1 Case 2 Case 3

Tavg (K) stage 1 2522.4 K 2574.2 K 2775 KTavg (K) stage 2 1817.7 K 1805.4 K 1601 KOverall FC conversion 85.5% 89.8% 83.1%SpecieCO 0.6489 0.5491 0.6266H2 0.2764 0.3290 0.2360CH4 0.0410 0.0445 0.0016CO2 0.0038 0.0416 0.0189H2O 0.0094 0.0106 0.0000O2 0.0000 0.0000 0.0000N2 0.0035 0.0076 0.0000H2S 0.0063 0.0072 0.0045Ar 0.0106 0.0104 0.0110

100 A. Slezak et al. / Powder Technology 203 (2010) 98–108

because commercial CFD software such as Fluent typically has thebuilt-in capability to allow use of a user-selected particle size pdfrepresented by several different diameter classes, and the individualparticle trajectories and burn out behaviors of each size class maythen be investigated individually. However, this built-in procedurerequires that all particles have a single diameter distribution, and alsohave the same density value and chemistry models. The study byCloke et al. [10] experimentally investigated the variation in coalproperties versus particle diameter for ten different world coals; thesecoals were selected to give a wide range of properties. The coals usedin this study were not density-separated. The experimental work byWu et al. [11] studied the composition of coarse and fine slags thatwere sampled for a Chinese commercial single-stage, downflow,slurry-fed entrained-flow gasifier unit. The coarse slag was collectedfrom the outlet of the lock hopper beneath the main gasifier unit,

Fig. 2. Computed gas phase temperature field: Fluent simu

while the fine slag was collected from the fine slag removal filter forthe water streams in the quench chamber and the downstream gasscrubber/convective cooler. Previous CFD simulation studies of coalgasification [3,7–9,12–27] are briefly summarized in Table 1; none ofthis previous work has simulated variations in particle densities.

Our present simulations have focused on the same gasifiergeometry, CFD mesh, and chemistry models as in Shi et al. [3], buthave utilized the ultimate and proximate analysis data for thePittsburgh #8 seam coal from the Bailey mine that is under study inthe current CPP, again at an operating pressure of 2.84 MPa. Theproximate and ultimate analysis data for the parent Bailey coal arepresented in Table 2. For the current simulations, the computedPittsburgh #8 coal Sauter-mean diameter of 108 μm and averagespecific gravity of 1.373 for the composite coal have been used; seeTable 3. For comparison purposes, the simulation by Shi et al. usingIllinois #6 coal assumed a specific gravity of 1.4 and an averageparticle diameter of 30 μm. (This average particle diameter is believedto be too small to represent conditions in a slurry-fed entrained-flowgasifier.)

For the present simulations using the Bailey coal, the same total coalmass flow rate as used by Shi et al. [3] has been used in both of thepresent simulations (28.04 kg/s), with the same split between the firstand second stages (78%/22%). However, because of the higher carboncontent of this Pittsburgh #8 coal relative to the Illinois #6 (86.08%versus 80.51%), the stage 1 inlet carrier gas (consisting of 94.4% O2) wasscaled up from 21.61 kg/s to 23.17 kg/s, and the slurry water flow rateswere scaled in the sameproportion (from11.62 kg/s to 12.31 kg/s). Thiswas done to maintain the same (O/C) ratio of 1.44, and (steam/C) ratioof 0.34 as was used by Shi et al. for Illinois #6 coal. Further details of thepresent case setup conditions and solution process have been given inthe thesis by Slezak [28].

3. Results

Details of the present Fluent CFD solutions using the DiscretePhase Method (DPM) for the commercial-scale two-stage upflowentrained-flow gasifier geometry previously studied by Shi et al. [3] atan operating pressure of 2.84 MPa are summarized and comparedherein. Additional comparisons have been presented by Slezak [28].Shi et al. [3] used the devolatilization chemistrymodel of Kobayashi etal. [4] and the char conversion models of Wen and Chuang [5]. TheCase 1 simulation is for the parent coal at a Sauter-mean diameter of108 μm and the mean density of 1373 kg/m3 (Table 3), while the Case2 simulation is for the injection of all 28 density/size cuts, each at the

lation of upflow entrained-flow gasifier; p=2.84 MPa.

Fig. 3. Computed gas velocity magnitude: Fluent simulation of upflow entrained-flow gasifier; p=2.84 MPa.

101A. Slezak et al. / Powder Technology 203 (2010) 98–108

appropriate injection mass flow rate (Table 3), and each with its ownvolatile chemical composition and proximate analysis [1,2]. A thirdsimulation (Case 3) has also been obtained using a devolatilizationmodel developed using secondary pyrolysis results from the PC CoalLab software developed by Niksa [6], and the char conversion modelsof Wen and Chaung. The gasifier geometry and the grid used in thesimulations are shown in Fig. 1, while the assumed model conditionsfor the Case 1, Case 2, and Case 3 runs are summarized in Table 4.

The computed average gas temperature in the first stage for Case 2is about 50 K hotter than results for Case 1 (Table 5), while the Case 2second stage average temperature is about 12 K cooler than for Case 1;these are modest differences. However, the computed exit gas speciesvolume fractions for these two solutions differ (Table 5). In particular,the exit CO volume percentage is 10% lower for the partitioned Case 2solution than for Case 1, while the exit H2 volume fraction increasedby over 5%. Computed exit volume fraction of CH4 increased by about0.5%, while the exit CO2 volume percent increased by about 4%. Theother gas species volume percentages are quite low, and are similarfor both solutions. The Case 3 solution (PC Coal Lab devolatilizationmodel) has exit gas volume fractions that compare more closely with

Fig. 4. Computed CO species mole fraction: Fluent simula

the Case 1 solution (Kobayashi devolatilization model, as used in Shiet al. [3]). However, the average temperatures differ dramaticallyfrom those of Case 1, being approximately 250 K hotter in stage 1 and215 K cooler in stage 2. This is believed to be due to the differentassumed volatile composition for the two models, where PC Coal Labpredicts that the predominant secondary pyrolysis product is soot,which was not produced at all in the Kobayashi model. Soot wasallowed to react in the gas phase with oxygen and rapidly consumesany available O2 in the first stage leading to higher temperatures, andgasifies with the available steam in the second stage. The overallpredicted fixed carbon conversion is 85.5% for Case 1, 90% for Case 2,and 83.1% for Case 3.

Average temperature contours for Cases 1 and 3 appear nearlysymmetric in the first stage, while for Case 2 there is an asymmetry inthe temperature distribution in the first stage of the gasifier (Fig. 2B).

This may be because the heaviest particles become trapped in thegasifier first stage. Mean velocity contours (Fig. 3) for all three caseslook more nearly the same than do the temperature contours,although asymmetry in the velocity distribution is visible in the firststage for Case 2 in Fig. 3B. The COmole fraction distributions for Case 1

tion of upflow entrained-flow gasifier; p=2.84 MPa.

Fig. 5. Computed H2 species mole fraction: Fluent simulation of upflow entrained-flow gasifier; p=2.84 MPa.

102 A. Slezak et al. / Powder Technology 203 (2010) 98–108

and Case 3 show higher concentrations in stage 2 as well as in thevicinity of the slag tap relative to levels for Case 2; compare Fig. 4A andC to Fig. 4B. This is consistent with the CO exit volume fractions shownin Table 5. Again, asymmetry is observed in the spatial contours in

Fig. 6. Fixed carbon conver

stage 1 for the Case 2 solution (Fig. 4B). Conversely, the H2 molefraction is higher in stage 2 for the Case 2 solution; see Fig. 5.

Predicted overall fixed carbon conversion is compared for the Case1 and Case 2 solutions for each particle size, for the SG1, SG2, and SG3

sion for Cases 1 and 2.

Fig. 7. Average particle residence time for Cases 1 and 2.

103A. Slezak et al. / Powder Technology 203 (2010) 98–108

density cuts in Fig. 6. Trends are quite similar for the two solutions,with nearly 100% conversion predicted for the three smallest particlediameter ranges (PS7, PS6, and PS5, or 50, 90, and 128 μm), and lowercarbon conversion predicted for the larger particle sizes. For example,between 60% (for SG3) and 40% (for SG1) conversion is predicted forthe largest particles (PS1, or 800 μm). Case 1 conversion levels werealways lower than those using the individual density fractions (Case

Fig. 8. Impact velocity normal to the wall versus axial location in second stage f

2) which is consistent with the lower temperatures for the averageddensity fractions used for Case 1.

The computed average particle residence times are compared forthese two solutions in Fig. 7. Residence times for the three smallestparticle size classes (up to 128 μm diameter) are approximately 3s,with residence times increasing for the larger particles, especially forthe Case 2 solution. There are significant differences in the predicted

or density class SG1 for Cases 1 and 2. Particles injected from stage 2 inlet.

Fig. 9. Impact velocity normal to the wall versus axial location in second stage for density class SG4 for Cases 1 and 2. Particles injected from stage 2 inlet.

104 A. Slezak et al. / Powder Technology 203 (2010) 98–108

residence times, with Case 2 consistently predicting larger values.Including accurate density variations, and the individual inclusion andtreatment of the larger size fractions (especially PS1, PS2, and PS3)while iterating the DPM solution for Case 2 results in the longerpredicted residence times, which are captured in the Case 1 solutiononly for the highest density and largest particle size fraction (SG4PS1;see Fig. 7D).

The computed average wall impact velocity distributions are quitesimilar for the Case 1 and Case 2 solutions; see Figs. 8 and 9 forexamples for particle injection from the stage 2 injection locationonly. Additional examples have been given by Slezak [28]. (Slurryinjection velocities were set to the same value of 50 m/s as used by Shiet al. [3] at all three injection locations.) Also, the wall impact velocityis largest near the stage 2 injection location, and then decreasessignificantly (to between 1 and 0.5 m/s). There also is an increase inimpact velocity near the exit, due to the contraction just prior to theexit. Impact velocities near the second stage injector are largest for thelarger particles (PS1; 800 μm). There is a trend for the heavierparticles to have higher wall impact velocities (SG4; Fig. 9). Also,impact velocities for second stage injection tend to be higher than forparticles that were injected into the gasifier first stage [28].

An example ofmeasured particle–wall impact probability results isshown in Fig. 10 along the length of the second stage. This depicts theprobability of particle–wall collision for injection of all 28 density/sizefractions into both the first and second stages assuming that all

Fig. 10. Impact probability versus second stage distance for 28 size/density fractions.Particles injected from both stage 1 and stage 2.

particles stick to the wall at their first impact with the wall. Impactprobability is highest near the second stage injection location. TheCase 1 solution predicts 10% of all particles never impact a wall, whileCase 2 predicts only 5% do not impact the wall. Again, further detailshave been given by Slezak [28].

Fig. 11 displays contours of computed impact particle mass flux forparticles injected from both stage 1 and stage 2 at the Sauter-meandiameter and composite average density of the Bailey coal. If oneassumes that 100% of the particles impact the wall and 80% of thecarbon has been converted prior to impact, then the averagemass fluximpacting the second stage wall from this rough estimate isapproximately equal to the area-averaged value obtained from thedata shown in Fig. 11. The mass flux is highest almost directly acrossfrom the second stage slurry injection port, as one would expect.

Significant differences have also been observed for the computedparticle trajectories, for different particle initial density and diameter.The dramatic differences in particle trajectories for the extreme endsof the particle size-density spectrum are highlighted in Fig. 12, wherethe computed particle trajectories until the initial particle impact withthe gasifier wall are shown for second stage injections of the lightest,smallest cut (SG1PS7) and the heaviest, largest cut (SG4PS1). Impactsfor the SG1PS7 cut appear to be distributed over the entire gasifiersecond stage wall, while the SG4PS1 particles all impact the wallalmost directly opposite the injection location. Also, the particle–wall

Fig. 11. Contours of mass flux (kg/m2s) for Sauter-mean diameter and compositeaverage density. Particles impact once and then stick.

Fig. 12. Particle trajectories until first wall impact for second stage injection of: A) cut SG1PS7 (SG=1.2, d=50 μm), and B) cut SG4PS1 (SG=3.3, d=800 μm).

105A. Slezak et al. / Powder Technology 203 (2010) 98–108

impact velocities are much larger for the SG4PS1 cut, being as large as10–14 m/s, as compared to average values of 1–3 m/s for the SG1PS7cut; see Figs. 8B and 9A. These trajectory and wall impact differenceswould significantly influence the location and rate of slag formation,which then would feed back into a secondary influence on the localgas phase properties (temperature; gas composition).

Grid independence checks for the present geometry have beenreported in the thesis by Slezak [28]. Simulation results for the setupsfor Case 1 and Case 2 were run to 10,000 iterations for increasing griddensities of 12,256 cells (the present mesh), 20,061 cells, 28,264 cells,and 40,112 cells. Centerplane contours of gas species mole fractions,average gas temperatures, and wall temperatures were presented bySlezak [28], along with the second stage particle–wall impact normalvelocities, and some localized visible differences were observed insome regions. However, no clear trends could be discerned for thesevariations versus the grid density. As an overview, the computed exitgas mole fractions are compared in Fig. 13 for Case 1 and in Fig. 14 forCase 2. Similar plots are given by Slezak [28] for the gasifier stage 1and stage 2 average temperatures. Only modest differences areobserved in these major gas species and average temperatures, so thepresent mesh (12,256 cells) is judged to be adequate to predict theseoverall gasifier characteristics. Also, particle–wall impact velocitiesgenerally did not change significantly as the grid density wasincreased. However, variations were observed in the predicted spatial

Fig. 13. Case 1 exit gas compositions for 12,256, 20,061, 28,264, and 40,112 gridelements.

contours of gas specie volume fraction, for some species and in somelocalized regions. So, it is concluded that the finer grid should be usedto obtain the most realistic overall simulations. However, this wouldresult in significant run time increases and was not deemed necessaryhere when looking at the particle–wall collisions.

Preliminary validation efforts have been summarized in the thesisby Slezak [28]. Radiant coal flow reactor (RCFR) experimentsconducted by SRI International as a part of the CPP for a pyrolysisrun, a combustion run, and a gasification run for a Bailey coal sample[29] were simulated in Fluent. Wen and Chaung's shrinking core charconversion model and the devolatilization model as implemented byShi et al. [3] were used, with the same chemical kinetic rateparameters as have been used in the present commercial-scalegasifier simulations. The computational model of the SRI RCFR isshown in Fig. 15. The SG2PS5 fraction was chosen as the coal density/size fraction for the validation study, because this fraction was thecenter point of the SRI test matrix.

The computational grid used (Fig. 15) contains 91,584 hexahedralcells; no grid sensitivity studies were conducted for this SRI PCFRgeometry The reactor is 1.2 m long, with a 12 mm ID. The coal wasinjected via the core flow inlet in Ar carrier gas, and any oxygen wasintroduced via the outer sheath flow. Steam was injected through theconical inlet shown in Fig. 15. Wall temperature boundary conditionswere chosen to match the experimental setup at SRI as closely aspossible. The inlet velocity of the coal particles was set at 24 cm/s. The

Fig. 14. Case 2 exit gas compositions for 12,256 and 40,112 elements.

Fig. 15. Computational model of SRI facility.

Fig. 17. Comparison of exit gas yields for SRI run 80.

Fig. 18. Comparison of exit gas yields for SRI run 81.

106 A. Slezak et al. / Powder Technology 203 (2010) 98–108

operating conditions for the three SRI runs chosen for this validationstudy are summarized in Table 6. Run 49 was a pyrolysis case, run 81was a combustion run, and run 80 was a gasification run at themaximum steam and O2 levels. The computed centerline gastemperatures for all three runs are shown in Fig. 16.

Results for the CO2, H2, and CO exit gas yields obtained from theCFD validation runs are compared to the SRI International experi-mental data for runs 80, and 81 in Figs. 17 and 18, respectively.

The CFD predictions of the CO and H2 yields for the pyrolysis case(run 49; not presented) did not agree with the experimental data.These discrepancies are either due to an experimental O2 leak thatwas discovered after the run was completed, or to the fact that tar andsoot were not included as pyrolysis species in the present devolati-lization model. It is noted that the volatile distribution predicted inthe CFD model [3] assumes that the product gases resulting frompyrolysis consist primarily of CH4, which is not found in the SRIpyrolysis results. The CO and H2 predictions are realistic for runs 80and 81, but CO2 is underpredicted for both of these cases (by 40% ofthe measured value for run 80 and 34% for run 81). Fixed carbonconversion predicted by Fluent for the gasification case (run 80) wasapproximately 72%, versus 88.3% measured by SRI. For run 81 thecomputed fixed carbon conversion was 47%, versus a measured levelof 72%.

Additional Fluent CFD simulations have been obtained for acommercial-scale single-stage downflow entrained-flow gasifier,

Table 6Operating conditions for SRI SG2PS5 density/size cut experimental runs [29].

Run # Oven length Coal loading Oxygen loading Steam loading SR(cm) (wt.%) (wt.%) (wt.%)

49 15 1.5 0.0 0.0 0.080 120 2.10 3.22 9.0 0.6881 120 1.88 2.88 0.0 0.68

Fig. 16. Computed centerline gas temperature for SRI validation runs.Fig. 19. Commercial-scale, single-stage, downflow, slurry-fed, entrained-flow gasifiergeometry.

Table 7Proximate and ultimate analyses of Kentucky #9 coal.

Component Kentucky #9 coal

Fixed carbon 41.99Volatiles 35.97Ash 8.84Moisture 13.2C 62.31H 4.20O 7.94N 1.35S 2.64Cl 0

Fig. 20. Comparison of exit gas composition for commercial-scale downflow gasifier.

107A. Slezak et al. / Powder Technology 203 (2010) 98–108

using the same chemistry models [4,5] and gasification UDF [3,28] aswere used to simulate the two-stage upflow gasifier. The geometryand 7058-cell axisymmetric grid used are shown in Fig. 19. Thepresent simulations were for a Kentucky #9 coal, to enablecomparison with published performance data for the commercialunit [30]. Proximate and ultimate analysis data for this coal are givenin Table 7.

Coal slurry flow rate (27.8 kg/s coal and 11.0 kg/s H2O), gas flowrate (22.5 kg/s of 96% O2), and operating pressure (2.69 MPa) wereselected to correspond to results given by Hornick and McDaniel [30].Coal particle average diameter and specific gravity were set tod=231 μm, SG=1.373, with a Rosin–Rammler size distributionbetween 25 μm and 850 μm.

The computed exit gas composition is compared with data fromHornick and McDaniel [30] in Fig. 20. Agreement between the CFDpredictions and measured data are reasonable, except for under-prediction of CO and overprediction of CO2 volume fractions. Thecomputed fixed carbon conversion for this simulation is 95%,

Fig. 21. Particle trajectories for injection of: A) cut SG1PS7 (SG=

compared to a measured value of 91%. Examples of the computedparticle trajectories are shown in Fig. 21 for the lightest, smallestparticles (SG1PS7; SG=1.2; d=50 μm) and the heaviest, largestparticles (SG4PS1; SG=3.3, d=800 μm). The lighter, smaller parti-cles tend to follow the gas flow and are recirculated in the gasifier,while the heavier, larger particles impact the wall nearer to the exitand do not recirculate. This is shown quantitatively in Fig. 22, wherethe percent of particles striking the wall is shown versus distancealong the gasifier for the two particle size/density cuts.

4. Conclusions

Fluent CFD simulations using DPM particle tracking have beenobtained for commercial-scale two-stage upflow and single-stagedownflow entrained-flow gasifier geometries, and the solutions havebeen used to examine variability in the computed gas phase averageexit conditions (temperature and species volume fractions), gas phaseflow field contours, particle char conversion, particle residence time,and particle–wall impact velocity for the 28 varying particle densityand size fractions under study in the current Coal Partitioning Project.For these density/size fractions, the density ranges between SG=1.2and SG=3.3, while particle diameter ranges between d=50 μm andd=800 μm.

For the two-stage upflow gasifier, the computed gas phase flowfield, computed particle char conversion, and computed residencetimes for some coal fractions, each differ significantly between thesimulation that used the parent coal properties (Case 1) and thesolution that used properties for each of the 28 different Bailey coaldensity/size fractions that have been separated and characterized inthe CPP project (Case 2). This indicates that in order to mostaccurately determine both the actual trajectories of coal particles foreach density and size fraction and the gasifier flow field, it is necessaryto also model the injections of each of these density/size cuts in theiterated DPM simulation. Significant differences have also beenobserved for the computed particle trajectories, as the particle initialdensity and diameter are varied. Solutions that were obtained usingtwo different devolatilization models generally compared reasonablywell.

Grid independence studies showed the grid used in the compar-ison studies was adequate for predicting exit gas composition andwall impact velocities. Preliminary validation studies using experi-mental data for the Bailey coal by SRI International in theirpressurized coal flow reactor at 30 atmospheres [29] indicatedadequate agreement for gasification and combustion runs, but pooragreement for a pyrolysis case. Additional simulation of a single-stagedownflow gasifier yielded an exit gas composition that was inreasonable agreement with published data.

1.2, d=50 μm), and B) cut SG4PS1 (SG=3.3, d=800 μm).

Fig. 22. Particle–wall impacts for lightest, smallest particles (SG1PS7; SG=1.2, d=50 μm) and heaviest, largest particles (SG4PS1; SG=3.3, d=800 μm).

108 A. Slezak et al. / Powder Technology 203 (2010) 98–108

Acknowledgements

This work has been supported under the DOE GasificationTechnology Program, funded by the National Energy TechnologyLaboratory in Morgantown, WV, as part of the University ResearchInitiative, through the “Collaboratory for Multiphase Flow Research,”RDS Contracts No. 41817M2318 and 41817M2100.

References

[1] N. Soundarrajan, M. Klima, S.V. Pisupati, Coal Partitioning Project MonthlyProgress Report, Project Progress Report, Pennsylvania State Univ, June 2007.

[2] N. Soundarrajan, M. Klima, S.V. Pisupati, Coal Partitioning Project MonthlyProgress Report, Project Progress Report, Pennsylvania State Univ, Jan 2008.

[3] S.P. Shi, S.E. Zitney, M. Shahnam, M. Syamlal, W.A. Rogers, Modelling coalgasification with CFD and discrete phase method, Journal of the Energy Institute79 (n4) (Dec 2006) 217–221.

[4] H. Kobayashi, J.B. Howard, A.F. Sarofim, Coal Devolatilization at High Tempera-tures, 16th Symp. (International) on Combustion, The Combustion Institute, 1976.

[5] C.Y. Wen, T.Z. Chaung, Entrainment coal gasification modeling, Ind. Eng. Chem.Process. Dev. 18 (1979) 684–695.

[6] S. Niksa, PC Coal Lab Version 4.0 User Guide and Tutorial, Niksa Energy Assoc. LLC,Belmont, CA, 2006.

[7] C. Chen, M. Horio, T. Kojima, Numerical simulation of entrained flow coal gasifiers.Part I: modeling of coal gasification in an entrained flow gasifier, ChemicalEngineering Science 55 (2000) 3861–3874.

[8] C. Chen, M. Horio, T. Kojima, Use of numerical modeling in the design and scale-upof entrained flow coal gasifiers, Fuel 80 (2001) 1513–1523.

[9] X.J. Liu, W.R. Zhang, T.J. Park, Modelling coal gasification in an entrained flowgasifier, Combustion Theory and Modelling 5 (2001) 595–608.

[10] M. Cloke, E. Lester, A. Belghazi, Characterization of the properties of size fractionsfrom ten world coals and their chars produced in a drop-tube furnace, Fuel 81(2002) 699–708.

[11] T. Wu, M. Gong, E. Lester, F. Wang, Z. Zhou, Z. Yu, Characterization of residualcarbon from entrained-bed coal water slurry gasifiers, Fuel 86 (2007) 972–982.

[12] Bouma, P.H., de Goey, L.P.H., Tummers, M.J., and Kiel, J.H.A. (1999). “NumericalModelling of an Entrained-Flow Gasification Simulator,” Computational Technol-ogies for Fluid / Thermal / Structural / Chemical Systems with IndustrialApplications—1999 (ASME Pressure Vessels and Piping Conf.), Boston, MA, Aug.1–5.

[13] C. Chen, T. Miyoshi, H. Kamiya, M. Horio, T. Kojima, On the scaling-up of a twostage air blown entrained flow coal gasifier, Canadian Journal of ChemicalEngineering 77 (1999) 745–750.

[14] M. Shahnam, M. Syamlal, D. Cicero, Numerical Modeling of Combustion andGasification Processes using the Discrete Particle Method, International JointPower Generation Conf, ASME Fuels & Combustion Technologies Div, MiamiBeach, FL, 2000, July 23–26.

[15] D.F. Fletcher, B.S. Haynes, F.C. Christo, S.D. Joseph, A CFD based combustion modelof an entrained flow biomass gasifier, Applied Mathematical Modelling 24 (n3)(March 2000) 165–182.

[16] C. Chen, M. Horio, T. Kojima, Numerical simulation of entrained flow coal gasifiers.Part II: effects of operating conditions on gasifier performance, ChemicalEngineering Science 55 (2000) 3875–3883.

[17] H. Tominaga, T. Yamashita, T. Ando, N. Asahiro, Simulator development ofentrained flow coal gasifiers at high temperature and high-pressure atmosphere,IFRF Combustion Journal, June 2002, Article No. 200004, http://www.industrial.combustion.ifrf.net/view_paper.html?paperId=14.

[18] M. Bockelie, M. Denison, Z. Chen, T. Linjewile, C. Senior, A. Sarofim, N. Holt, CFDModeling for Entrained Flow Gasifiers, Gasification Technologies Conference, SanFrancisco, CA, Oct. 27–30 2002.

[19] H. Liu, T. Kojima, Theoretical study of coal gasification in a 50 ton/day HYCOLentrained flow gasifier. I. Effects of coal properties and implications, Energy andFuels 18 (4) (July/Aug 2004) 908–912.

[20] H. Liu, T. Kojima, Theoretical study of coal gasification in a 50 ton/day HYCOLentrained flow gasifier. II. Effects of operating conditions and comparison withpilot-scale experiments, Energy and Fuels 18 (4) (July/Aug 2004) 913–917.

[21] J.B. Roucis, J.M. Kline, J.D. Smith, Advanced Modeling of Gasifiers: GasificationProcess Modeling using Computational Fluid Dynamics (CFD), GasificationTechnologies Conference, Washington, DC, Oct. 6 2004.

[22] H. Watanabe, M. Otaka, J. Inumaru, Development of numerical simulationtechnique for coal gasification on entrained flow coal gasifier (modeling of coalgasification reaction and prediction of gasifier performance), Transactions of theJapan Society of Mechanical Engineers, Part B 70 (n695) (July 2004) 1856–1863.

[23] C. Guenther, R. Breault, The Effect of Gas–Solids Dispersion on a Two Fluid Modelof a Transport Gasifier, AIChE Annual Meeting and Fall Showcase, Cincinnati, OH,Oct. 30–Nov. 4 2005.

[24] H. Tominaga, Y. Matsushita, H. Aoki, T. Yamashita, T. Akiyama, T. Miura, Thedevelopment of entrained flow coal gasification simulator, Nihon EnerugiGakkaishi/Journal of the Japan Institute of Energy 85 (n1) (Jan. 2006) 36–41.

[25] J. Lu, L. Yu, X. Zhang, S. Zhang, W. Dai, Hydrogen production from a fluidized-bedcoal gasifier with in situ fixation of CO2 Part I: numerical model, ChemicalEngineering & Technology 31 (2008) 197–207.

[26] M. Forstner, G. Hofmeister, M. Jooler, J. Dahl, M. Braun, S. Kleditzsch, R. Scharler, I.Obernberger, CFD simulation of ash deposit formation in fixed bed biomassfurnaces and boilers, Progress in Computational Fluid Dynamics 6 (2006)248–261.

[27] N. Syred, K. Kuriawan, T. Griffiths, R. Ray, Development of fragmentation modelsfor solid fuel combustion and gasification as subroutines for inclusion in CFDcodes, Fuel 86 (2007) 2221–2231.

[28] Slezak, A. A. (2008). “Modeling of Particle Trajectories of Coal Size and DensityFractions in a Gasifier,” M. S. Thesis, West Virginia Univ., Dept. of Mechanical andAerospace Engineering, Morgantown, WV; URL: https://eidr.wvu.edu/etd/doc-umentdata.eTD?documentid=6016.

[29] D.J. Eckstrom, R. Malhotra, Coal Partitioning High Pressure Gasification Experi-ments, Project Progress Report, SRI International, Mar 2008.

[30] Hornick, M. J. and McDaniel, J. E. (2002). “Tampa Electric Polk Power StationIntegrated Gasification Combined Cycle Project,” Final Technical Report, U.S. DOECoop. Agrmt.DE-FC-21-91MC27363.


Recommended