Anda di halaman 1dari 13

Energy Fuels 2010, 24, 21332145

Published on Web 02/22/2010

: DOI:10.1021/ef901497b

Numerical Comparison of the Drag Models of Granular Flows


Applied to the Fast Pyrolysis of Biomass
K. Papadikis, S. Gu,*, A. Fivga, and A. V. Bridgwater

School of Engineering Sciences, University of Southampton, Highfield, Southampton, SO17 1BJ, United Kingdom, and

School of Engineering and Applied Science, Aston University, Aston Triangle, Birmingham B4 7ET, United Kingdom
Received December 8, 2009. Revised Manuscript Received February 9, 2010

The paper presents a comparison between the different drag models for granular flows developed in the
literature and the effect of each one of them on the fast pyrolysis of wood. The process takes place on an
100 g/h lab scale bubbling fluidized bed reactor located at Aston University. FLUENT 6.3 is used as the
modeling framework of the fluidized bed hydrodynamics, while the fast pyrolysis of the discrete wood
particles is incorporated as an external user defined function (UDF) hooked to FLUENTs main code
structure. Three different drag models for granular flows are compared, namely the Gidaspow, Syamlal
OBrien, and Wen-Yu, already incorporated in FLUENTs main code, and their impact on particle
trajectory, heat transfer, degradation rate, product yields, and char residence time is quantified. The
Eulerian approach is used to model the bubbling behavior of the sand, which is treated as a continuum.
Biomass reaction kinetics is modeled according to the literature using a two-stage, semiglobal model that
takes into account secondary reactions.

reported in the literature6,9-14 to account for the momentum


exchange between the gas and solid phases and have been
applied to dispersed two-phase flow simulations with great
success. Most of these models have been studied and compared
with experimental data in several cases, such as the studies of
Kafui et al.,4 Li and Kuipers,5 and Taghipour et al.15 The work
of Kafui et al.,4 concerned the simulation of pressure drop-gas
superficial velocity using two models (pressure gradient force
model, fluid density based buoyancy model) in an EulerianLagrangian framework, using the continuous single-function
correlation of Di Felice13 to account for the drag forces in dense
and dilute regimes. The study showed significant differences in
the pressure drop-superficial gas velocity profiles in the fixed
bed regime and corresponding significant differences in the
prediction of minimum fluidization velocity. The work of
Li and Kuipers5 examined the occurrence of heterogeneous
flow structures in gas-particle flows and the impact of nonlinear
drag force for both ideal particles (elastic collision, without
interparticle friction) and nonideal particles (inelastic collision,
with interparticle friction). Several drag models were compared,10,13,14,16 and the study showed that heterogeneous flow
structures exist in systems with both nonideal and ideal
particle-particle interaction. For the ideal particle interaction it was shown that heterogeneous flow structures are
caused purely by the nonlinearity of the gas drag and that the
stronger the dependence of drag to voidage, the more
heterogeneous flow structures develop. The weak dependence of voidage to drag causes more homogeneous flow

1. Introduction
Fluidized beds have been the center of modeling attention
for many years, and various simulation approaches have
been developed, ranging from Lattice-Boltzmann methods
(LBM)1 to discrete particle (DPMs, Eulerian-Lagrangian)2-5
and two-fluid models (TFMs, Eulerian-Eulerian).6,7 More
recently, discrete bubble models (DBMs)8 have been developed and been applied to fluidized beds, increasing the
potential of industrial scale simulations beyond the capabilities of TFMs.
The Eulerian formulation of the granular medium, using
the kinetic theory of granular flows, has made the realization
of fluidized bed simulations, less computationally intensive.
The particulate phase is treated as a continuum with an
effective viscosity, thus the method is also called a two-fluid
approach. The calculation of the momentum exchange coefficient of gas-solid systems is very different in comparison of a
single sphere surrounded by a fluid. When a single particle
moves in a dispersed two-phase flow, the drag force is affected
by the surrounding particles. Several correlations have been
*To whom correspondence should be addressed. Telephone: 023 8059
8520. Fax: 023 8059 3230. E-mail: s.gu@soton.ac.uk.
(1) Ladd, A. J. C.; Veberg, R. J. Stat. Phys. 2001, 104, 1191.
(2) Feng, Y. Q.; Xu, B. H.; Zhang, S. J.; Yu, A. B.; Zulli, P. AIChE. J.
2004, 50 (8), 17131728.
(3) Hoomans, B. P. B.; Kuipers, J. A. M.; Mohd Salleh, M. A.; Stein,
M.; Seville, J. P. K. Powder Technol. 2001, 116 (2-3), 166177.
(4) Kafui, K. D.; Thornton, C.; Adams, M. J. Chem. Eng. Sci. 2002,
57, 23952410.
(5) Li, J.; Kuipers, J. A. M. Chem. Eng. Sci. 2003, 58, 711718.
(6) Gidaspow, D. Multiphase Flow and Fluidization: Continuum and
Kinetic Theory Descriptions; Academic Press: New York, 1994.
(7) Enwald, H.; Peirano, E.; Almstedt, A. E.; Leckner, B. Chem. Eng.
Sci. 1999, 54, 311328.
(8) Bokkers, G. A.; Laverman, J. A.; Van Sint Annaland, M.;
Kuipers, J. A. M. Chem. Eng. Sci. 2006, 61, 55905602.
(9) Syamlal, M.; OBrien, T. J. AIChE. Symp. Ser. 1989, 85, 2231.
(10) Wen, C.-Y.; Yu, Y. H. Chem. Eng. Prog. Symp. Ser. 1966, 62,
100111.
r 2010 American Chemical Society

(11) McKeen, T.; Pugsley, T. Powder Technol. 2003, 129, 139152.


(12) Gibilaro, L. G.; Di Felice, R.; Waldram, S. P.; Foscolo, P. U.
Chem. Eng. Sci. 1985, 40, 18171823.
(13) Di Felice, R. Int. J. Multiphase Flow 1994, 20, 153159.
(14) Happel, J. AIChE. J. 1958, 4, 197.
(15) Taghipour, F.; Ellis, N.; Wong, C. Chem. Eng. Sci. 2005, 60,
68576867.
(16) Hill, R. The effects of fluid inertia on flow in porous media. Ph.D.
Dissertation; Cornell University: 2001.

2133

pubs.acs.org/EF

Energy Fuels 2010, 24, 21332145

: DOI:10.1021/ef901497b

Papadikis et al.

structures. The simulations were also carried out in an


Eulerian-Lagrangian framework. The work of Taghipour
et al.15 concerned the Eulerian-Eulerian simulation of a
fluidized bed comparing the effect of the drag correlations of
Gidaspow,6 Syamlal-OBrien,9 and Wen and Yu10 at the
pressure drop and sand bed expansion. All studies showed
good qualitative gas-solid flow patterns.
Fluidized beds are the most widely used type of reactor for
fast pyrolysis, as they offer a number of advantages, such as
high heat transfer rates and good temperature control. Several
models have been developed and applied to biomass pyrolysis.17-30 Most of the models apply numerical analysis to
determine heat, mass, and momentum transport effects in a
single biomass particle varying the pyrolysis conditions.
Usually, the influence of parameters such as, size, shape,
moisture, reaction mechanisms, heat transfer rates, and particle shrinkage is the main objective of study. The numerical
investigations of di Blasi.17-22 are typical examples of single
particle models that study the heat, mass, and momentum
transport through biomass particles by varying the pyrolysis
conditions, feedstock properties (implementing different chemical kinetics mechanisms), particle sizes, effect of shrinkage,
etc. The models provide very useful information about different
conditions of pyrolysis and product yields and can be an excellent guide for every researcher on the field. Saastamoinen,23
and Saastamoinen and Richard24 studied the effect of drying
to devolatilization stating that the surface temperature of the
particles after drying can exceed those of the initiation of
devolatilization, meaning that drying and pyrolysis may overlap. The models of Babu and Chaurasia25-30 treat the pyrolysis process in a similar way as the one of di Blasi and focus
more on the heat transfer effects and estimation of optimum
parameters for biomass pyrolysis. They also study the effect
of the enthalpy of reaction on product yields and vary
the shrinkage parameters of biomass close to the complete
degradation of the particles. They found that their approach
gave better validation with experimental data compared to
those of di Blasi.
The main research of the authors has been focused on the
incorporation of the single particle models presented in the

literature into a unified computational model that simulates


the complete process of fast pyrolysis inside a bubbling
fluidized bed.31-37 In this way, several important parameters,
such as heat transfer coefficients from the bubbling bed, char
and pyrolysis vapor residence times, biomass particle trajectories, and effect of particle size and shape, can be monitored
and quantified, something that is impossible by single particle
models alone. This unified CFD model can greatly aid the
design and optimization of pyrolysis equipment by specifying
the operation limits of the fluidized bed reactor regarding the
type, the size, and the shape of biomass particles, as well as the
high heat transfer regimes for more efficient biomass degradation and consequently efficient char elutriation at the freeboard.
The scope of the current paper is to quantify the effect of
the various drag models for granular flows that have been
presented in the literature, to the fast pyrolysis of biomass.
The different flow pattern and bubble distribution that will be
created by these models will have a certain impact on the heat
transfer coefficient, the momentum transport from the bed to
the biomass particle, the final product yields, and more
importantly the char and vapor residence times in the reactor.
The reaction kinetics is based on a two-stage, semiglobal
mechanism, with kinetic constants suitable for wood pyrolysis
according to Chan et al.,38 for the primary stage and Liden
et al.,39 and di Blasi17 for the secondary reactions. This scheme
has been chosen for this study because it can predict the
correct behavior of wood pyrolysis including the dependence
of the product yields on temperature.18,22,38 The simulation
results are compared with experimental data, whenever it is
possible, that comes from the pyrolysis of spruce in the same
reactor.
2. Model Description
2.1. Experiment. The biomass fast pyrolysis experiment
was carried out in Aston Universitys fluidized bed reactor.
The reactor has a capacity of 100 g/h, is constructed of
stainless steel, and is 40 mm in diameter and 260 mm high.
The fluidizing gas was nitrogen, whereas sand was used as a
fluidizing and heat transfer medium with an average particle
diameter of 440 m. The flow rate of nitrogen used was
sufficient to provide 3 times the minimum fluidizing velocity
(MFV) to the bed. The residence time of the pyrolysis vapors
in the reactor was approximately 0.93s at a reaction temperature of 538 C. The main apparatus of the fast pyrolysis
unit consists of the feeder, the reactor, and the liquid product
collection system and is shown in Figure 1.
2.2. Numerical Model. The dimensions of the 100 g/h fast
pyrolysis reactor are illustrated in Figure 2. Nitrogen flows

(17) Di Blasi, C. Combust. Sci. Technol. 1993a, 90, 315339.


(18) Di Blasi, C. Chem. Eng. Sci. 1996, 51, 11211132.
(19) Di Blasi, C. Chem. Eng. Sci. 2000, 55, 59996013.
(20) Di Blasi, C. Biomass Bioenerg. 1994, 7, 8798.
(21) Di Blasi, C. Ind. Eng. Chem. Res. 1996b, 35, 3746.
(22) Di Blasi, C. J. Anal. Appl. Pyrol. 1998, 47, 4364.
(23) Saastamoinen, J. Model for drying and pyrolysis in an updraft
gasifier. In Advances in Thermochemical Biomass Conversion; Bridgwater,
A. V. Ed.; Blackie: London, 1993; pp 186-200.
(24) Saastamoinen, J.; Richard, J.-R. Combust. Flame 1996, 106, 288
300.
(25) Babu, B. V.; Chaurasia, A. S. Modelling and Simulation of
Pyrolysis: Influence of Particle Size and Temperature. Proceedings of
the International Conference on Multimedia and Design, 2002a, 4, Mumbai,
India; pp 103-128.
(26) Babu, B. V.; Chaurasia, A. S. Modelling and Simulation of
Pyrolysis: Effect of Convective Heat Transfer and Orders of Reactions.
Proceedings of International Symposium and 55th Annual Session of IIChE
(CHEMCON-2002), 2002b, OU, Hyderabad, India; pp 105-106.
(27) Babu, B. V.; Chaurasia, A. S. Energ. Convers. Manage. 2003a, 44,
22512275.
(28) Babu, B. V.; Chaurasia, A. S. Energ. Convers. Manage. 2003b, 44,
21352158.
(29) Babu, B. V.; Chaurasia, A. S. Modelling and Simulation of
Pyrolysis of Biomass: Effect of Heat of Reaction. Proceedings of International Symposium on Process Systems Engineering and Control (ISPSEC
03) for Productivity Enhancement Through Design and Optimisation 2003c,
IIT-Bombay, Mumbai, India; pp 181-186.
(30) Babu, B. V.; Chaurasia, A. S. Energ. Convers. Manage. 2004c, 45,
12971327.

(31) Papadikis, K.; Bridgwater, A. V.; Gu, S. Chem. Eng. Sci. 2008, 63
(16), 4218-4227.
(32) Papadikis, K.; Gu, S.; Bridgwater, A. V. Chem. Eng. Sci. 2009, 64
(5), 1036-1045.
(33) Papadikis, K.; Gerhauser, H.; Bridgwater, A. V.; Gu, S. Biomass
Bioenerg. 2009, 33 (1), 97-107.
(34) Papadikis, K.; Gu, S.; Bridgwater, A. V.; Gerhauser, H. Fuel
Process. Technol. 2009, 90 (4), 504-512.
(35) Papadikis, K.; Gu, S.; Bridgwater, A. V. Chem. Eng. J. 2009, 149
(1-3), 417427.
(36) Papadikis, K.; Gu, S.; Bridgwater, A. V. Biomass Bioenerg. 2010,
34 (1), 2129.
(37) Papadikis, K.; Gu, S.; Bridgwater, A. V. Fuel Process. Technol.
2010, 91 (1), 6879.
(38) Chan, W. R.; Kelbon, M.; Krieger, B. B. Fuel 1985, 64, 1505
1513.
(39) Liden, A. G.; Berruti, F.; Scott, D. S. Chem. Eng. Commun. 1988,
65, 207221.

2134

Energy Fuels 2010, 24, 21332145

: DOI:10.1021/ef901497b

Papadikis et al.

size of the particles used in the experiments. Bigger rigs and


commercial plants use much larger particles in the range of
2-5 mm.
The different bubble distribution that will occur due to the
different granular flow drag models will have an impact on
the heat transfer conditions and consequently will affect the
degradation rate of the particles and subsequently the final
product yields and the residence time of char and vapors. The
numerical model will try to identify any significant differences between the models and compare the results with reallife experiments. The simulations last until the particles are
entrained from the reactor independently of the simulation
time needed to achieve that.
3. Mathematical Model
3.1. Multiphase Flow Governing Equations. The simulations of the bubbling behavior of the fluidized bed were
performed by solving the equations of motion of a multifluid
system. An Eulerian model for the mass and momentum for
the gas (nitrogen) and fluid phases, was applied, while the
kinetic theory of granular flow was applied for the conservation of the solids fluctuation energy. The Eulerian model is
already incorporated in the main code of FLUENT, and its
governing equations are expressed in the following form.
Mass Conservation. Eulerian-Eulerian continuum modeling is the most commonly used approach for fluidized bed
simulations. The accumulation of mass in each phase is
balanced by the convective mass fluxes. The phases are able
to interpenetrate, and the sum of all volume fractions in each
computational cell is unity.
Gas Phase.

Figure 1. Experimental apparatus. Adapted from Coulson M,


EC Project Bioenergy Chains Final Report, Aston University,
ENK6-2001-00524.

g Fg
r 3 g Fg g 0
t

s Fs
r 3 s Fs s 0
t

Solid Phase.

Momentum Conservation. Newtons second law of motion


states that the change in momentum equals the sum of forces
on the domain. In gas-solid fluidized beds the sum of forces
consists of the viscous force r 3 Cs , the solids pressure force
rps, the body force sFsg, the static pressure force s 3 rp and
the interphase force Kgs(ug - us) for the coupling of gas and
solid momentum equations by drag forces.
Gas Phase.

Figure 2. Fluidized bed reactor geometry used in the simulation.

through a porous plate at the bottom of the reactor at a


velocity of U0 = 0.6 m/s and temperature of 303 K. The
superficial velocity is approximately 3 times greater than the
minimum fluidizing velocity Umf of the reactor, which is
typically around 0.2 m/s using a sand bed with average
particle diameter of 440 m.40 The sand bed has been
preheated to 815 K as it was also performed in the experimental procedure.
The biomass particle of density of 700 kg/m3 is injected at
the center of the sand bed, which has been previously
fluidized for 1 s. Momentum is transferred from the bubbling
bed to the biomass particle as well as from the formed
bubbles inside the bed. The studied biomass particle was
chosen to be 350 m in diameter, which is more-or-less the

g Fg g
r 3 g Fg g Xg
t

-g 3 rp r 3 g g Fg g Kgs ug -us

Solid Phase.
s Fs s

r 3 s Fs vs Xs -s 3 rp -rps r 3 s
t
s Fs g Kgs ug -us
4
where the solid-phase stress tensor is given by,

(40) Geldart, D. Powder Technol. 1973, 7, 285.

2135


s s rus ruTs s

2
s - s r 3 us I s
3

Energy Fuels 2010, 24, 21332145

: DOI:10.1021/ef901497b

Papadikis et al.

motion. This phenomenon is quantified by the term called


frictional viscosity

The Gidaspow interphase exchange coefficient is,


3 s g Fg jus -ug j -2:65
Kgs Cd
g
for g > 0:8
ds
4
Kgs 150

2s g
s Fg jus -ug j
1:75
for g e 0:8
2
ds
g ds

s, fr

24
1 0:15g Res 0:687 
g Res

and
ds Fg jus -ug j
Res
g

The Syamlal-OBrien kinetic viscosity is


p 

s Fs ds s
2
s, kin
 1 s g0, ss 1 ess 3ess -1
63 -ess
5

The Syamlal-OBrien interphase exchange coefficient is,


!
3s g Fg
Res
jus -ug j
Cd
10
Kgs 2
4 ur, s ds
ur, s

22
The solids pressure ps, which represents the normal force
due to particle interactions is given by
ps s Fs s 2Fs 1 ess 2s g0, ss s

where

Cd

4:8
0:63 p
Res =ur, s

and
ur, s 0:5A -0:06Res
q
0:06Res 2 0:12Res 2B -A A2

12

with
13

A g4:14 , B 2:65
g , for g > 0:85

14

where Cs is defined in eq 5. The diffusion coefficient of


granular temperature ks according to eq 6 is given by:

The Wen-Yu interphase exchange coefficient is,

ks
15

where
Cd

24
1 0:15g Res 0:687 
g Res

25

26

the collision dissipation energy as


s

The tangential forces due to particle interactions are summarized in the term called solids shear viscosity, and it is defined as

where the collision viscosity of the solids s,col is


r
4
s
s, col s Fs ds g0, ss 1 ess
5

p 
2
150Fs ds s
6
1 s g0, ss 1 ess
3841 ess g0, ss
5
r
s
2
2Fs ds s g0, ss 1 ess

The radial distribution function g0,ss is defined as


2
!1=3 3 -1

s
5
g0, ss 41 s, max

16

The bulk viscosity s is a measure of the resistance of a


fluid to compression, which is described with the help of the
kinetic theory of granular flows
r
4
s
17
s s Fs ds g0, ss 1 ess
3

s s, col s, kin s, fr

-ps I s s : r 3 us r 3 ks 3 r 3 s -s gs 24

or

s g Fg jus -ug j -2:65


3
Kgs Cd
g
ds
4

23

Fluctuation Energy Conservation of Solid Particles.


The solid phase models discussed above are based on
two crucial properties, namely the radial distribution function g0,ss and granular temperature s. The radial distribution function is a measure for the probability of interparticle
contact. The granular temperature represents the energy
associated with the fluctuating velocity of particles.


3
s Fs s r 3 s Fs us s
2 t

11

A g4:14 , B 0:8g1:28 , for g e 0:85

20

which contains the angle of internal friction . This expression of frictional viscosity is valid if the volume fraction of
solids is constant, which is reasonable in the dense packed
bed state of particles. The Gidaspow6 kinetic viscosity is
p

2
10Fs ds s
4
 1 s g 0, ss 1ess
21
s, kin
96s g0, ss 1 ess
5

where the drag coefficient is given by


Cd

ps sin
p
2 I2D

121 -e2ss g0, ss


p
Fs 2s 3=2
s
ds

27

and the transfer of kinetic energy gs as

18

jgs -3Kgs s

28

19
(41) Ranz, W. E.; Marshall, W. R. Chem. Eng. Prog. 1952a, 48, 141
146.
(42) Ranz, W. E.; Marshall, W. R. Chem. Eng. Prog. 1952a, 48, 173
180.

At the packed state of the bed, the stresses are dominated


by interparticle friction rather than collisions and fluctuating
2136

Energy Fuels 2010, 24, 21332145

: DOI:10.1021/ef901497b

Papadikis et al.

An analytical discussion of the solid-phase properties can


be found in ref 43.
3 2. Forces on Discrete Particles. The coupling between the
continuous and discrete phases has been developed in a UDF
to take into account the bubbling behavior of the bed. For an
analytical discussion of this section the reader is referred to
the previous work done by the authors in this aspect.31
Assuming a spherical droplet with material density of Fd
inside a fluid, the rate of change of its velocity can be
expressed as44
!
dud
f
Fcon
Fm
ucon -ud g 1 29
u
dt
Fd

Figure 3. Two-stage, semiglobal model.49

and the drag force between solid and liquid is computed for a
modified particle volume fraction p
d
p
37
s d
and an effective continuum viscosity eff,con


p -1:55
eff , con 1 dm

where f is the drag factor and u is the velocity response time


Fd dd
18con
2

30

If the volume fraction of the space among the solid


particles, if they were closely packed, is larger than the liquid
fraction

There are several correlations for the drag factor f in the


literature.45-47 The one used in this study is the correlation of
Putnam47

s > s

31

f 0:0183Rer for 1000 e Rer < 3  105

32

dg d 1 -s =s

3.3. Reaction Kinetics. The reaction kinetics of biomass


pyrolysis is modeled using a two-stage, semiglobal model.49
The mechanism is illustrated in Figure 3.
The mechanism utilizes the Arrhenius equation which is
defined as
42
Ki Ai exp -Ei =RT

According to Kolev,48 if bubble three-phase flow (i.e., solid


particles in bubbly flow) is defined, two subcases are distinguished. If the volume fraction of the space among the solid
particles, if they were closely packed, is smaller than the
liquid fraction (in this case the Eulerian sand)

The values of the kinetic parameters were obtained by Chan


et al.38 for the primary pyrolysis products, while the fourth
and fifth reactions were obtained from Liden et al.39 and Di
Blasi,17 respectively. The choice of a different set of parameters used in the simulation was mainly done to predict as
correctly as possible the fast pyrolysis of wood. The kinetic
parameters of the three-step model of Chan et al.38 were used
for the primary products because, when coupled with secondary reaction kinetics, they can predict, at least qualitatively, the correct behavior of wood pyrolysis.22 However,
the secondary reaction kinetic constants were taken from the
work of Liden et al.39 and di Blasi17 because these studies
included secondary reaction effects that depend on the concentration of tar vapors from flash pyrolysis in the porous
matrix of the solid, something that is more suitable for the
current study. The elemental analysis of the pine wood used
in the experiments of Chan et al. can be found in refs 38
and 50.
The model associates, via a UDF, the reaction kinetics
mechanism with the discrete biomass particle injected in the
fluidized bed. The particles properties change according to
the reaction mechanism due to the phase transition phenomena. Momentum and heat transfer on the particle are

34

where


1 -dm
d
dm

40

are surrounded by gas and the drag force can be calculated


between one single solid particle and gas as for a mixture
dg
p
41
g dg

The second term on the right-hand side of the equation


represents the gravity and boyancy force, while the third
term represents the unsteady force of virtual mass, which is
expressed as


F Vd ducon dud
33
Fm con
2
dt
dt

39

then only

Rer 2=3
for Rer < 1000
f 1
6

s < s

38

35

then the theoretical possibility exists that the particles are


carried only by the liquid. The hypothesis is supported if we
consider the ratio of the free setting velocity in gas and liquid
s
Fd -Fg Fs
wdg

.1
36
wds
Fd -Fs Fg
Due to great differences between gas and liquid densities,
the particles sink much faster in gas than in a liquid. Therefore, the drag force between gas and solid particles is zero,
(43) Boemer, A.; Qi, H.; Renz, U. Int. J. Multiphas. Flow 1997, 23 (5),
927944.
(44) Crowe, C. T.; Sommerfeld, M.; Tsuji, Y. Multiphase Flows with
Droplets and Particles; CRC Press LLC: 1998.
(45) Schiller, L.; Naumann, A. Ver. Deut. Ing. 1933, 77, 318.
(46) Clift, R.; Gauvin, W. H. Proc. Chemeca 1970, 1, 14.
(47) Putnam, A. ARS JNl. 1961, 31, 1467.
(48) Kolev, N. I. Multiphase Flow Dynamics 2. Thermal and Mechanical Interactions, 2nd ed.; Springer: 2005.

(49) Shafizadeh, F.; Chin, P. P. S. ACS Symp. Ser. 1977, 43, 57-81.
(50) Chan, W. R.; Kelbon, M.; Krieger, B. B. Ind. Eng. Chem. Res.
1988, 27, 22612275.

2137

Energy Fuels 2010, 24, 21332145

: DOI:10.1021/ef901497b

Papadikis et al.
Table 1. Simulation Parameters

property

value

biomass density, Fw
biomass particle diameter, dp
biomass specific heat capacity, Cpw
char specific heat capacity, Cpc
biomass thermal conductivity, kw
char thermal conductivity, kc
superficial velocity, U0
gas density, Fg
gas viscosity, g
gas specific heat capacity, Cp,g
gas thermal conductivity, kg
solids particle density, Fs
sand specific heat capacity, Cp,s
sand thermal conductivity, ks
mean solids particle diameter, ds
restitution coefficient, ess
initial solids packing, s
static bed height
bed width
heat of reaction

700 kg/m
350 m
1500 J/(kg K)
1100 J/kg K
0.105 W/(m K)
0.071 W/(m K)
0.6 m/s
1.138 kg/m3
1.663  10-5 kg/(m s)
1040.67 J/(kg K)
0.0242 W/(m K)
2500 kg/m3
835 J/(kg K)
0.35 W/(m K)
440 m
0.9
0.6
0.08 m
0.04 m
H = -255 kJ/kg

Nu 2 0:9Rep, mf 0:62 dp =db 0:2

Rep, mf

45

The effective thermal conductivity keff and effective specific heat capacity Cp,eff are given by

Cp eff Cp c jCp w -Cp c jw

47

The heat transfer coefficient is evaluated from the wellknown Ranz-Marshall41,42 correlation, when the particle is
carried only by the fluidizing gas
hdp
2:0 0:6Rep 1=2 Pr1=3
kg

50

For the implementation of the model, certain parameters


have been quantified and assumptions have been made in
order to provide, as much as possible, an insight to the fast
pyrolysis process in bubbling beds.
(1) The particles used in the simulation were assumed to be
totally spherical, whereas the particles used in experiments can
be found on all sorts of shapes. The actual sphericity of the
particles greatly differs from 1. This would have an impact on
the drag and virtual mass forces and consequently on the
trajectory of the particle inside the reactor.
(2) The model assumes a plug flow profile at the inlet of the
reactor.
(3) The geometry of the reactor has been discretized using a
structured grid. The average side length of the computational
cells is about 1 mm, resulting in a total number of 10 440 cells.
(4) The time-step used for the simulation was on the order
of 10-5 s. The reason for this small time advancements was
that the computational algorithms in the UDF had to converge simultaneously with the algorithms of FLUENT. The
explicit central difference scheme used for the discretization of
the heat diffusion equation demands very small time-steps due
to the small radial discretization resulting from particles of
350 m in diameter. For further analysis of the CFD code
structure together with the flow-chart and advantages and
drawbacks of the model, the reader is referred to the previously published work of the authors in this aspect.32
The simulation parameters are shown in Table 1.

r 0

46

Fg Umf dp
g

4. Model Parameters

r R

keff kc jkw -kc jw

49

where

The boundary condition at the surface of the particle is


given by

T 
hT -Ts -keff 
44
r 
and at the center of the particle

T 
0

r 

wood
fixed
wood
char
wood
char
3Umf
nitrogen (303 K)
nitrogen (303 K)
nitrogen (303 K)
nitrogen (303 K)
sand
fixed
fixed
uniform distribution
value in literature
fixed value
fixed value
fixed value
Koufopanos et al.52

coefficient according to the findings of Collier et al.,51

calculated according to the new condition in the UDF, and


the variables associated with it are updated in each time-step.
Intraparticle secondary reactions due to the catalytic effect
of char are taken into account, resulting in secondary vapor
cracking.
3.4. Heat Transfer. The heat conduction along the radius
of the particle is calculated by solving the heat diffusion
equation for an isotropic particle. The particle is discretized
in the UDF for the complete pyrolysis model and the reader
is referred to the work previously done by the authors in this
aspect.32





1
F
2 T
FCp eff T 2
keff r
-H 43
t
r r
r
t

Nu

comment

48

However, since the diameter of the biomass particle is


smaller than the diameter of the sand particles, a modified Nusselt number is used to calculate the heat transfer

5. Results
5.1. Fluidized Bed Hydrodynamics and Particle Trajectories.
Figures 4-6 illustrate the fluidized bed hydrodynamics

(51) Collier, A. P.; Hayhurst, A. N.; Richardson, J. L.; Scott, S. A.


Chem. Eng. Sci. 2009, 59, 46134620.

(52) Koufopanos, C. A.; Papayannakos, N.; Maschio, G.; Lucchesi,


A. Can. J. Chem. Eng. 1991, 69, 907915.

2138

Energy Fuels 2010, 24, 21332145

: DOI:10.1021/ef901497b

Papadikis et al.

Figure 4. Volume fraction of sand with instantaneous position of


the biomass particle, Gidaspow drag model.

Figure 6. Volume fraction of sand with instantaneous position of


the biomass particle, Wen-Yu drag model.

mainly caused by the tendency of the particles to move


toward the wall of the reactor as it is shown in Figure 7
where gas velocities tend to be much lower than at the central
axis of the reactor. This is valid for all the drag models
compared in this study, however the bubble distribution
created from the Gidaspow drag model moved the particle
to the right-hand side of the reactor in contrast to the
Syamlal-OBrien and Wen-Yu drag models, which moved
the particle on the left-hand side. The 2D nature of the
simulation excludes the mixing of the biomass particles in
the z-direction of the reactor, however the same behavior was
also observed in the previous papers by the authors,35,37
where 3D geometry was considered. The inclusion of the
pyrolysis vapors mass source in the simulation greatly discourage the use of a 3D geometry since the simulation times
approach the order of months even with the use of 16
processors. Another phenomenon that is of great importance
is the transverse mixing of the particles. Despite the dominance of the motion in the y-direction of the reactor, we can
easily notice that the particles have moved across the whole
width of the bed (x-direction) before their entrainment. This
comes in great agreement with the experimental study of
Hoomans et al.,3 where they showed the validity of this
phenomenon using the positron emission particle tracking
(PEPT) technique, and the numerical study of Xu and Yu.53
Despite the fact that these studies were performed using 2D
geometries, the same phenomenon was also observed in
the previous numerical investigations performed by the
authors31,35,37 where a 3D geometry was used.
5.2. Heat Transfer. Figure 8 shows the temperature rise of
the particles for the different drag models. The temperature
rise in the Gidaspow and Syamlal-OBrien model appears to
be relatively similar, however a higher temperature rate is
observed for the Wen-Yu drag model, represented by a
steeper surface temperature profile at the first 0.1 s after the

Figure 5. Volume fraction of sand with instantaneous position of


the biomass particle, Syamlal-OBrien drag model.

together with the instantaneous position of the biomass


particle indicated by the black spot. The snapshots were
taken for 1 s time intervals, from the time of injection at 1 s
until close to the entrainment time of approximately 4 s.
Figure 7 illustrates the complete trajectory of the particles for
the three different drag models. The residence time of the
particles was finally determined as 2.92, 3.3, and 2.6 s for the
Gidaspow, Syamlal-OBrien, and Wen-Yu drag models,
respectively.
One can observe that the residence time of the particles is
somewhat higher, though not unreasonable, compared to
what the fast pyrolysis process suggests, i.e., < 2 s. This is

(53) Xu, B. H.; Yu, A. B. Chem. Eng. Sci. 1997, 52, 27852809.

2139

Energy Fuels 2010, 24, 21332145

: DOI:10.1021/ef901497b

Papadikis et al.

Figure 7. Particle trajectories inside the fluidized bed reactor.

transfer between a fluidized bed of particles larger than the


injected particles. In our case the bed consists of particles
of 440 m in diameter, whereas the biomass particles are
350 m in diameter. Collier et al.51 showed that when a
particle finds itself among fluidized particles of larger diameter, the heat transfer coefficient appears to have a constant value depending on the ratio of the diameter of the
particles and the minimum fluidization velocity of the bed.
This is because the conductive effect of the larger particles to
the smaller ones is negligible and convective effects are
dominant. This approach is used in the current study. Therefore, whenever the heat transfer coefficient appears to have
this constant value, it means that the particle finds itself in a
densely packed zone of sand particles. The other values of the
heat transfer coefficient represent the motion of the particle
inside a bubble, or the splash zone, or the freeboard of the
reactor. This gives a good indication of the region that the
particle is instantly located.
The high heat transfer values, close to 700-750 W/(m2 K),
that are observed in all of the three diagrams clearly indicate
the motion of the particle close to the wall of the reactor,
where the wall heat transfer effects become more noticeable.
In the current simulation only the convective effects from the
wall were taken into account and the collision of the particles
with the wall were neglected, since the actual contact time of
the particles with the wall is extremely small. However,
Figure 9 in conjunction with Figure 7 clearly show the wall
effect on the rise of the heat transfer coefficient. This
becomes even more noticeable when the particle flies toward
the outlet of the reactor and the heat transfer coefficient
drops significantly in all cases, only to rise again in the actual
outlet of the reactor where the x-velocity of the gases increase
significantly due to the small diameter of the outlet.

Figure 8. Surface and center temperatures of the particles.

particle injection. In the case of the Wen-Yu drag model, the


particle reaches the splash zone of the reactor much faster
than in the Gidaspow and Syamlal-OBrien drag models. In
this region, high velocity gradients are dominant due to the
eruption of bubbles, and the slip velocity between the
fluidizing gas and the particles is high, resulting in a very
high convective heat transfer coefficient. This can be easily
seen in Figure 9, where the heat transfer coefficient becomes
very large for the case of the Wen-Yu model almost
instantly. In the other two cases a certain amount of time
(0.3 s) is needed for the particles to experience the same
high heat transfer rates.
The constant value of the heat transfer coefficients, which
is close to 320 W/(m2 K), comes from the correlation of
Collier et al. in eq 49, which is appropriate for the heat
2140

Energy Fuels 2010, 24, 21332145

: DOI:10.1021/ef901497b

Papadikis et al.

Figure 9. Heat transfer coefficient.

appears to have more intense y-velocity fluctuations, and this


is because, as it was discussed earlier, it reaches the splash
zone of the reactor almost 0.2 s faster than the other particles.
This high y-velocity fluctuation gave rise to the very high
heat transfer coefficients that were observed earlier and
consequently the more rapid density drop of the particle.
Comparing Figures 8, 9, and 11. one can conclude that the
unstable splash zone of the reactor is the most desirable
region for fast pyrolysis. In other words, the violent velocity
gradients that occur due to bubble eruption and sand
recirculation have a higher impact on the particles that float
on the top of the bed, in contrast with the particles that are
immersed deeply inside it. Also, the high velocity gradients
give rise to the virtual mass force that can greatly contribute
to the violent ejection of the particles from the sand bed to the
freeboard of the reactor, thus making the elutriation phenomenon easier, when the density drop of the particles has
reached the desirable level.
It has to be noted that for the current simulation the
particles were considered to be totally spherical, something
that differs from reality. Also, shrinking and attrition effects
were assumed not to play an important role and therefore
were neglected in the simulation. For the complete analysis
of the momentum transport model, the reader is referred to
the previous study of the authors,31 and for the modeling of
the impact of shrinkage to ref 35.
5.4. Product Yields. Figure 12 shows the final product
yields (vapors, gases, char) percentage by weight of original
wood. We can observe that the different drag models actually
have an impact on the final product yield of condensable
volatiles (vapors), char, and gases, while there is still a small
percentage of wood that has not yet reacted. The final
product yields for the three cases are given in Table 2, and
the experimental data is given in Table 3.
In general terms, we can observe that there is a similarity
between the final vapors and the gases in the experiment and
the simulation. However there is a significant difference on

Figure 10. Density drop of the particles.

The effect of the heating rates can be seen in Figure 10,


where the density drop due to the reaction mechanism is
illustrated. The Gidaspow and the Syamlal-OBrien models
appear to have the same effect on the density drop of the
particle, whereas the Wen-Yu model has a more rapid effect
due to the higher heat transfer rates that appear during the
simulation. From Figure 10 one can also observe the onset of
pyrolysis, which comes close to 0.2 s, where, according to
Figure 8, the temperature is close to 750 K. The density drop
of the particles has a major effect on their trajectory inside
the reactor since their motion is heavily dependent upon their
density and consequently their mass that is involved on the
particle motion equations given in Section 3.2.
5.3. Particle Dynamics. Figure 11 shows the instantaneous
velocity components of the particles inside the reactor until
their entrainment time. All the drag models appear to have a
similar effect on the average velocity magnitude of the
particles with some exceptions regarding their instantaneous
positions. For example, the particle on the Wen-Yu case
2141

Energy Fuels 2010, 24, 21332145

: DOI:10.1021/ef901497b

Papadikis et al.

Figure 11. Particle velocities inside the fluidized bed reactor.

Figure 12. Product yields.


Table 2. Product Yields from the Simulation

Gidaspow
Syamlal-OBrien
Wen-Yu

Table 3. Experimental data and product yields

wood

vapors

char

gases

pyrolysis
temperature C

3.5%
2.0%
6.4%

62%
63%
60%

20.5%
20.8%
20.0%

14.0%
14.2%
13.6%

525
525
528

the final char content, and differences of up to 7.5% can be


observed. This is mainly due to the reaction kinetics mechanism and not to the fluidized bed operation. Due to the lack
of analytic kinetic data regarding spruce fast pyrolysis, the
selection of kinetic data that would as closely as possible
2142

feedstock
moisture content (wt %)
ash content (wt %, dry basis)
pyrolysis temperature
vapor residence time

spruce
6.96
0.32
538 C
<0.93s

product yields (wt %, dry basis)


char
gases
vapors
organics
closure (wt %, dry basis)

12.72
15.58
68.65
56.22
96.65

Energy Fuels 2010, 24, 21332145

: DOI:10.1021/ef901497b

Papadikis et al.

Figure 13. Volume fraction of pyrolysis vapors, Gidaspow drag model.

Figure 14. Volume fraction of pyrolysis vapors, Syamlal-OBrien drag model.

literature that make use the same reaction mechanism such


as the work of di Blasi.22 It has to be also noted that the
specific chemical kinetics represent the degradation of large
pine wood particles where differences on the percentages of

match the behavior of a softwood like this was a necessity.


The kinetic data of Chan et al.38 refer to the pyrolysis of pine
wood, which belongs to the same softwood family. The
simulation results highly agree with data published in the
2143

Energy Fuels 2010, 24, 21332145

: DOI:10.1021/ef901497b

Papadikis et al.

Figure 15. Volume fraction of pyrolysis vapors, Wen-Yu drag model.

cellulose, hemicellulose, and lignin as well as the different


amount and composition of inorganic matter lead to different primary yields.18 Also, the heating rates that are commonly used in the laboratory for the determination of the
reaction kinetics of biomass cannot match the high heating
rates that are provided by a fluidized bed.
The impact of the different drag is mainly noticeable on
the final amount of wood that has not reacted. For the cases
of Gidaspow and Syamlal-OBrien drag models, the difference in the final percentage of wood is relatively minor,
whereas in the case of the Wen-Yu drag model the difference can be up to almost 4%. This is explained by the faster
entrainment of the biomass particle, which in the case of
Wen-Yu and Syamlal-OBrien models there is a difference
of approximately 0.7 s. However, the higher residence time of
the char particle inside the reactor is not a desirable phenomenon for fast pyrolysis due to the catalytic effect of char,
aiding the secondary cracking of the vapors. It has to be
noted that the extraparticle secondary cracking of the vapors
is not taken into account in the current simulation and only
the intraparticle phenomena are considered.
Figures 13-15 illustrate the evolution of volatiles in the
reactor during the pyrolysis of the biomass particle. The
actual residence time of the vapors comes close to 0.8 s,
something that is difficult to notice from the contour plots
due to scale limitations. This phenomenon comes in perfect
agreement with the experimental data where the vapor
residence time was observed to be less than 0.93 s. The
different drag models, however, have a major effect on the
flow field of the vapors due to the different bubble distribution and consequently the fluidizing gas flow field that
interacts with the produced vapors. In the cases of the
Gidaspow and Syamlal-OBrien drag models there is still
a larger amount of vapors inside the reactor (Figures 13
and 14), whereas in the case of the Wen-Yu drag model the

amount of vapors left is approximately 1 order of magnitude


lower. As was mentioned earlier, this is due to the faster
entrainment of the biomass particles that consequently
stopped the production of vapors inside the reactor. The
simulation of the vapor flow field inside the reactor is one of
the most important parameters in the operation of the
fluidized bed reactors for fast pyrolysis since it greatly
influences the final bio-oil yields.
6. Conclusions
The paper presented a comparison of the different drag
models for granular flows available in the literature and their
effect on the fast pyrolysis of biomass. Several conclusions can
be made regarding the complete fast pyrolysis process.
One can easily state that the different drag models certainly
have an impact on the trajectory of the biomass particle, the
heat transfer coefficient, and consequently on the degradation, the char, and the vapor residence time. The results
showed that in the case of the Wen-Yu drag model, the heat
transfer coefficient reached a maximum value of approximately 780 W/(m2 K) compared to the models of Gidaspow
and Syamlal-OBrien, where the maximum values reached
750 and 650 W/(m2 K), respectively. This was mainly observed
because of the fast movement of the particle toward the splash
zone of the reactor in the case of the Gidaspow and Wen-Yu
model. The residence time of the particle was greatly affected
by this effect, since in that case the residence time was reduced
by a maximum of 0.7 s when comparing Wen-Yu (3.6 s) and
Syamlal-OBrien (4.3 s) models. In the case of the Gidaspow
model the residence time was determined at 3.9 s. This period
of time is of great importance in fast pyrolysis, where the
optimum residence time is less than 2 s. It was also shown that
the trajectory of the particle was different in all three cases,
however in all of them the particle moved toward the wall of
2144

Energy Fuels 2010, 24, 21332145

: DOI:10.1021/ef901497b

Papadikis et al.

m = mass, kg
Nu = Nusselt number
p = pressure, Pa
Pr = Prandtl number
R = universal gas constant, J/molK
Re = Reynolds number, dimensionless
t = time, s
T = temperature, K
U0 = superficial gas velocity, m/s
ui = velocity, m/s
V = volume, kg/m3
wi = free settling velocity, m/s

the reactor due to the dominant formation of the gas bubbles at


the central axis of the reactor. This phenomenon acts as a decelerating factor in the entrainment of the particle since the fluidizing
gas velocities are much smaller close to the walls of the reactor.
In the case of the final product yields, we can see that the
residence time of the particle is the factor that mainly defines
the outcome. The difference in the final char yield between the
experiments and the simulation is mainly due to the reaction
kinetics mechanism and the differences between the two types
of wood (pine wood for the reaction kinetics and spruce in the
experimental procedure). However, there is a definite similarity at the final closure of the system, where in almost all cases
the total conversion exceeded 95%, something that indicates
that the simulations are quite realistic in conjunction with the
very representative residence times of char and vapors. Quantitatively, the Gidaspow model predicted 62% vapors, 20.5%
char, and 14.0% gas occurring at a pyrolysis temperature of
525 C; in the Syamlal-OBrien model, 63% vapors, 20.8%
char, and 14.2% gas occurring at a pyrolysis temperature of
525 C; and in the Wen-Yu model, 60% vapors, 20.0% char,
and 13.6% gas occurring at a pyrolysis temperature of 528 C.
The results were comparable to the experimental data, which
produced 68.65% vapors, 12.72% char, and 15.58% gas
occurring at a pyrolysis temperature of 538 C.
Despite the fact that significant differences between the drag
models were observed, it is still very difficult to say which one
gave the better prediction for various reasons mainly concerning the running time limitations of the model and the postprocessing of the data. However, it was shown that the drag
model has a definite impact in the simulation of fast pyrolysis in
fluidized beds, which can be quantified and examined.

Greek Letters
s = collision dissipation of energy, kg/s3m
H = heat of reaction, J/kg
i = volume fraction, dimensionless
i = granular temperature, m2/s2
i = bulk viscosity, kg/sm
Fi = shear viscosity, kg/ms
Fi = density, kg/m3
= velocity response time, s
= stress tensor, Pa
gs = transfer rate of kinetic energy, kg/ms3
= mass fraction
Subscripts
b = bed material
c = char
con = continuous phase
col = collision
d = droplet
D = Drag
dm = disperse phase maximum packing
eff = effective
fr = frictional
g = gas
i = general index
k = radial position
kin = kinetic
m = mixture
mf = minimum fluidization
p = particle
r = relative
s = solids
T = stress tensor
= velocity
m = virtual mass
w = wood

Nomenclature
Ai = pre-exponential factor, 1/s
Cd = drag coefficient, dimensionless
Cp = specific heat capacity, J/kgK
di = diameter, m
E = activation energy, J/mol
ess = restitution coefficient, dimensionless
g = gravitational acceleration, m/s2
g0,ss = radial distribution coefficient, dimensionless
h = convective heat transfer coefficient, W/m2K
I = stress tensor, dimensionless
I2D = second invariant of the deviatoric stress tensor,
dimensionless
f = drag factor, dimensionless
Fi = force, N/kg
k = thermal conductivity, W/mK
ks = diffusion coefficient for granular energy, kg/sm
Kgs = gas/solid momentum exchange coefficient, dimensionless

2145

Anda mungkin juga menyukai