Вы находитесь на странице: 1из 12

Article

Cite This: Energy Fuels 2018, 32, 10683−10694 pubs.acs.org/EF

Integrated Particle- and Reactor-Scale Simulation of Pine Pyrolysis


in a Fluidized Bed
M. Brennan Pecha,† Emilio Ramirez,‡ Gavin M. Wiggins,‡ Daniel Carpenter,† Branden Kappes,§
Stuart Daw,‡ and Peter N. Ciesielski*,†

Biosciences Center, National Renewable Energy Laboratory, 1503 Denver West Parkway, Golden, Colorado 80401, United States

Oak Ridge National Laboratory, 2360 Cherahala Boulevard, Knoxville, Tennessee 37932, United States
§
Department of Mechanical Engineering, Colorado School of Mines, 1500 Illinois Street, Golden, Colorado 80401, United States
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: We report results from a multiscale computational modeling study of biomass fast pyrolysis in an experimental
laboratory reactor that combined the hydrodynamics predicted by a two-fluid model (TFM) with predictions from a finite
Downloaded via UNIV OF PRETORIA on September 3, 2019 at 09:59:16 (UTC).

element method (FEM) simulation of heat and mass transfer and chemical reactions within biomass particles. The experimental
pyrolyzer consisted of a 2 in. (5.1 cm) diameter bubbling fluidized bed reactor (FBR) fed with milled pine pellets. The
predicted FBR hydrodynamics included estimates of the residence times that the gas and biomass particles spend in the reactor
before they exit. A single-particle FEM simulation was constructed on the basis of the geometry and heat transfer properties
determined from optical and X-ray computed tomography measurements of wood and char particles collected from the
experimental FBR, along with previously proposed pyrolysis reaction kinetics. Taken together, the combined TFM and FEM
simulation results predicted net bio-oil yields at the reactor exit that agree well with experimental observations, without any
arbitrary fitting parameters. The combined computational models also provided practical information about the most important
reactor and feedstock parameters.

1. INTRODUCTION devoid of spatial parameters) have also been developed.10,11


Today, pyrolysis is recognized in the renewable energy sector Global kinetic modeling is the simplest to perform because of
as a viable pathway for renewable biofuels, chemicals, and high- the lack of coupling to transport phenomena, which are critical
value carbonaceous materials.1 Although the technology is very in determining process outcomes in many pyrolysis scenarios.
popular, it is difficult to implement at industrial scales as a Coupling models of transport phenomena to detailed chemical
result of the complex, heterogeneous nature of biomass reaction mechanisms often require large computational
feedstocks and multiphase reactor operations.2 Nearly every resources, which can surpass capabilities of current day high-
aspect of pyrolysis impacts the product yield temperature, heat performance computing resources, although some progress has
transfer mechanism, collection system, biomass composition been made recently with lumped-species mechanisms and
and microstructure, ash content, particle size, and others.3 The particle-scale physical models.12−14
goal of fast pyrolysis is to produce bio-oil via rapid heat transfer 1.2. Simulating Realistic Particle Geometries. Single-
(>100 °C s−1) to biomass particles (such as wood) in the particle models have emerged as a useful approach for studying
presence of an inert carrier gas, which quickly removes feedstock-specific effects and other physical phenomena
condensable vapors and non-condensable product gases. Once commonly observed in the laboratory. Given sufficient
condensed, the vapors produce high yields of liquefied bio-oil. feedstock characterization data, such as particle size, shape,
Many different types of biomass pyrolysis reactors have been composition, density, and porosity, such models may be
proposed and tested over the years, but one of the most parametrized for specific biomass feedstocks. This facilitates a
common reactor designs used in laboratory studies is the more accurate representation of the inherent biocomplexity
fluidized bed reactor (FBR), in which biomass particles are and heterogeneity of real biomass feedstocks and can thereby
injected into a preheated bed of denser particles (such as sand) account for species-specific variations in conversion out-
that are “fluidized” by an upward flowing gas into a state that is comes.15 In addition, modeling pyrolysis at the single-particle
similar to a boiling liquid.4 To determine optimal reactor level allows for insights into the interconnected phenomena of
configurations and operating conditions, many researchers heat transfer, mass transfer, and conversion kinetics in the
have developed reasonable models to predict heat transfer, context of physically accurate geometries using well-studied
chemical reactions, and mass transfer for this type of reactor,5,6 engineering approaches.
but often experimental correlations are used in scaleup as a One of the most important findings from laboratory studies
result of the complexity and unreliability of these models. of biomass fast pyrolysis is that the highest yields of bio-oil are
1.1. Challenges to Modeling Biomass Pyrolysis achieved when the feed particle size is less than 0.5 mm.16 For
Phenomena. There are essentially three length scales at
which pyrolysis models have been developed: the reactor Received: July 3, 2018
scale,7 the single-particle scale,8 and the chemical species level.9 Revised: August 17, 2018
Global or semi-global scale models (i.e., empirical kinetics Published: September 4, 2018

© 2018 American Chemical Society 10683 DOI: 10.1021/acs.energyfuels.8b02309


Energy Fuels 2018, 32, 10683−10694
Energy & Fuels Article

each reactor, there is an optimal particle size that achieves the the motion of all of the individual particles in FBRs as a result
highest yields within the respective operating envelope. It has of limits of current computers and the high computational cost
been proposed that this trend is due to the faster heat and mass of tracking the thousands or millions of particles present in
transfer rates that can be achieved with the relatively large typical reactors. Thus, more approximate methods, such as the
surface area/volume ratios of the smaller biomass particles.17,18 two-fluid modeling (TFM) approach, have been pursued as a
Similarly, as a result of the anisotropic exterior geometry, way to reduce computational overhead.40 A recent TFM
which is typical of biomass particles, higher particle aspect modeling work illustrated the importance of using multiple
ratios (D/L) provide a higher yield of bio-oil.19 particle sizes for accurate pyrolysis conversion times, although
Recent technological advances have allowed for the use of X- it did not account for intraparticle gradients.41 In addition,
ray computed tomography (XCT) to characterize micrometer- recent reviews have detailed efforts to advance pyrolysis reactor
scale macroporosity in materials such as wood and char. While simulations,42,43 specifically methods for linking the physics
the technique has been used for coal as early as 1996 (as and chemistry at different spatial and temporal scales to
reviewed by Mathews et al.20), only recently has it been used achieve the correct global reactor performance.44 However,
for biomass. The development of XCT, starting in the early one remaining topic that has received little attention is the
1970s,21 did not evolve to the low-micrometer resolution until mixing behavior of non-spherical biomass particles and their
the late 1980s, when the use of the current scintillation-type residence times in FBRs.
detector was coupled to a charge-coupled device to achieve In this work, we present a multiscale integration approach
resolutions of 2.5 μm.22,23 Improvements in computing power, similar to that previously employed by Wurzenburger and
lab source brilliance, and data collection techniques have colleagues,45 wherein reactor- and particle-scale modeling
steadily improved the resolution of lab-scale XCT to 50 nm approaches are combined to provide an approximate
(e.g., Zeiss Ultra series X-ray microscopes) and 350 nm (e.g., description of the key physics and chemistry while reducing
Zeiss Versa series X-ray microscopes). Hyväluoma et al.24 used the computational overhead. In this work, TFM predictions of
XCT to study porosity characteristics of various biochars, the hydrodynamics were used to estimate particle and gas
finding that the pore size did not change dramatically with the residence times, which were then combined with single-particle
temperature but that the original properties of the biomass simulations of realistic biomass particle evolution during
were strong predictors of the properties of the char. Similar pyrolysis to estimate the overall yield of bio-oil for typical
findings were reported by Jones et al.25 XCT of chars from laboratory FBR conditions. We expect that this type of
gasification by Watanabe26 showed that they preserved the approach to integrated, multiscale modeling represents an
same pore structure as the biomass starting material and that exemplar computational framework that will enable many types
this anisotropy forced reaction progression to primarily occur of rapid screening simulations to estimate and interpret the
along the direction of the grains rather than perpendicular to effects of key experimental pyrolyzer parameters, such as
them. Thus, it is critical to represent the anisotropic porosity geometry, carrier gas flow rate, temperature, and residence
and permeability in pyrolysis simulations to realistically times, and realistic biomass feedstock parameters, such as
represent experimental observations. species and size−shape distributions.
Biomass particles typically exhibit anisotropic exterior
geometries. In one early study, Di Blasi implemented a 2. EXPERIMENTAL SECTION
model that includes particle multidimensional shrinking, 2.1. Fluid Bed Pyrolyzer. The experimental 2 in. fluid bed reactor
convection, conduction, and radiation, multistep kinetics, and (2FBR) at the National Renewable Energy Laboratory (NREL) was
moisture evaporation.27 However, the kinetic scheme was not used as the basis for the computational simulations of biomass fast
well-verified. Bellais developed a cylinder model for slow pyrolysis studied here. The key relevant features of this reactor are
pyrolysis with secondary reactions but did not include gas illustrated in Figure 1. While the 2FBR is part of a more extensive
experimental research reactor system that includes two FBRs, an
species.28 Another work studied the effect of the particle size initial FBR for pyrolysis and a second FBR for catalytic vapor-phase
and aspect ratio on the yields of products from pyrolysis, upgrading, the focus of this paper is on the first reactor (pyrolyzer) .
although it did not include water evaporation or mass transfer In this reactor, biomass particles and fluidizing gas are fed into a
by diffusion, and particle shrinking is not included.19 Paulsen et bubbling fluidized bed of hot sand, and the wood is thermally
al. developed an interesting model to track carbohydrate decomposed into gas, tar, and char products, which exit from the top
degradation inside the particle but used an isotropic model, of the reactor. Figure 1 provides general information about the reactor
ignored secondary reactions, and did not include diffusive mass geometry and typical operating conditions. More information about
transfer.29 Some studies even hypothesized a liquid-phase the reactor system is available in the online documentation at https://
ccpcode.github.io/docs-2fbr/ and the study by Howe et al.46
pyrolysis intermediate that bubbled as it vaporized.30−33
In the experiments of interest here, the biomass under study was
1.3. Multiscale Computational Fluid Dynamics (CFD) pine wood that was initially pelletized and then crushed and sieved
Modeling Approach. Fluid bed pyrolyzers are commonly through a 2 mm screen before being fed into the FBR. Each pyrolysis
used in laboratory and industrial settings to take advantage of experiment was run at a reactor temperature of 500 °C, with the wood
the constant flow of gas over all particles, which allows for high feed rate set at 1 kg/h for 2.5 h to achieve stable operation, so that
heat transfer from gas to solids.4 The gas and solids mixing and steady-state mass balances could be made. The initial sand charge
contacting patterns in FBRs are extremely complex. CFD remained in the reactor throughout each experimental run.
provides a way to simulate the spatiotemporal details about 2.2. Biomass Feedstock Properties. Statistical properties for the
pressures, velocities, flows, and concentrations that are either external size and shape of the as-fed biomass particles and char
particles collected from the FBR experiments were optically
impossible or extremely difficult to obtain experimentally.34 determined by MicroTrac using their PartAn3D analyzer, which
CFD has also been used in numerous studies of gas−solid makes three-dimensional (3D) images of particle samples and
fluidized beds,35 but a very limited number of biomass fast tabulates relevant parameters for up to hundreds of thousands of
pyrolysis CFD studies have addressed hydrodynamic ef- particles.47 The data provided by MicroTrac was processed and
fects.7,36−39 It is generally not feasible to directly simulate binned in MATLAB to summarize statistics that could be used to

10684 DOI: 10.1021/acs.energyfuels.8b02309


Energy Fuels 2018, 32, 10683−10694
Energy & Fuels Article

Table 1. Key Input Parameters for Two-Fluid Reactor-Scale


Simulation in MFIX
property value unit
particle diameter (sand) 500 × 10−6 m
particle density (sand) 2500 kg/m3
particle density (wood) 500 kg/m3
particle density (char) 80 kg/m3
temperature 773 K
pressure (inlet) 133 kPa
fluidizing N2 (range) 0.249 m/s
minimum fluidization 0.0565 m/s
coefficient of restitution 0.9
angle of repose 55 deg
friction coefficient 0.1

gas and solid phases for mass, momentum, and species in the fluid
domain are respectively given by

(εgρg ) + ∇(εgρg ug) = 0
∂t (1)

(εsρs ) + ∇(εsρs us) = 0
∂t (2)

(εgρg ug ) + ∇(εgρg ug ug ) = ∇τg − εg∇Pg + εgρg g − β(ug − us, m)
Figure 1. (Left) Schematic illustration of geometry and operating ∂
conditions of the 2 in. FBR system for thermochemical conversion of (3)
biomass. As biomass particles are conveyed through the pyrolyzer and
elutriate from the top, they are converted into light gases, bio-oil, and ∂
(εs, mρs, m us, m) + ∇(εs, mρs, m us, mus, m)
char to varying degrees depending upon their residence times. The ∂t
sand particles comprising the bulk of the fluidized bed are large and = ∇τs, m − εs, m∇Pg + εs, mρs, m g − β(ug − us, m)
dense enough to remain in the reactor. (Right) Example of the (4)
conceptual control volume and boundary conditions described below ∂
for use in the 2D axisymmetric single-particle simulations. (εgρg Xg, n) + ∇(εgρg Xg, nug ) = ∇Dg, n∇Xg, n
∂t (5)
construct representative “nominal” particles in the particle-scale ∂
(εs, mρs, m Xs, m , n) + ∇(εs, mρs, m Xs, m , nus, m) = ∇Ds, m , n − ∇Xs, m , n
simulations described below. ∂t
Internal microstructures of the biomass feed particles were (6)
characterized with X-ray micro-computed tomography (CT) using a
where g, s, s,m, g,n, and s,m,n describe gas, solid, solid-phase m, gas-
ZEISS Versa 520 X-ray microscope (XRM) or X-ray computed
phase species n, and solid-phase m species n, respectively, ε is the
tomography (XCT) operating at 30 kV and 2 W (67 μA) using the
volume fraction, ρ is the density, u is the velocity vector, τ is the stress
4× lens and no filter. This produced 3D CT images of the internal
tensor, P is the pressure, X is the mass fraction, D is the diffusion
structure with a voxel size of 3.4 μm, such as those depicted in Figure
coefficient, and β is the coefficient for the interphase force between
5, by convolving the information from 1600 to 2400 two-dimensional
solid and gas phases. The gas−solid momentum transfer, in the
(2D) X-ray images as particle samples were rotated through 360°.
interphase force, used the Syamlal−O’Brien49 correlation for the drag
model. To model solid transport properties, such as solid pressure and
3. COMPUTATIONAL APPROACH AND viscosity, the kinetic theory of granular flow,48 together with the
METHODOLOGY Schaeffer frictional stress tensor formulation50 and the sigmoidal
3.1. Reactor Hydrodynamic Simulations. The global 3D blending stress51 function, relates the computed solid temperature
hydrodynamics of the NREL 2FBR pyrolyzer were simulated with with solid transport properties. Furthermore, the numerical simulation
multiphase flow with interphase exchanges (MFIX),48 which is an used the species equation for tracking purposes, and there were no
open source CFD package specifically developed for modeling chemical reactions. Separate mass inlet boundary conditions were
multiphase reacting flows. In this case, the specific objective was to applied for the gas distributor and biomass inlet. A pressure outlet
use the two-fluid approximation to estimate the residence times of boundary condition defined the outlet at the top of the reactor, and
biomass and char particles. Values of the MFIX input parameters used wall boundary conditions were treated as no slip.
for these simulations are summarized in Table 1. The Sauter mean MFIX is capable of simulating extremely complex systems of
diameters calculated from the 3D particle Feret lengths were 58, 278, multiphase reacting flows. The specific conditions in the 2FBR
344, 426, and 543 × 10−4 cm. When simulations were run with a pyrolyzer make some important simplifications possible that reduce
density of 500 kg/m3, the fresh wood particles were found to have an the computational demand. Specifically, the 2FBR is operated such
infinite residence time (they elutriated to the bottom of the reactor) that temperature gradients are minimized and the large excess of
until they convert to char with a density of 80 kg/m3. Thus, running a diluent fluidizing gas relative to the biomass vapors makes it possible
larger wood density would not change the effective particle residence to neglect volume changes in gas flow through the reactor.
time, which was determined using eq 18, where Uwood is 0 and Uchar Biomass and char residence time distributions and average
was calculated to have a known value so that the results from initial velocities in the reactor were determined in this study by seeding
density of 500 kg/m3 were used. the simulated biomass inflow stream with tracer particles having the
The two-fluid approach approximates the gas and solid phases as density of either raw biomass or devolatilized char. Runs were
interpenetrating continua. The conservation equations solved for the completed using five different particle sizes (at a single fluidizing gas

10685 DOI: 10.1021/acs.energyfuels.8b02309


Energy Fuels 2018, 32, 10683−10694
Energy & Fuels Article

ij ∂u u yz
ρjjjj + (u∇) zzzz = −∇ρ
flow) to determine how the average particle residence times were
j ∂t εp z
k {
affected by the particle size. To estimate drag properties, the biomass

l
oji |
zyo
and char particles were treated as having spherical shapes

ojjμ(∇u + (∇u) ) − 3 μ(∇u)Izz}


+ ∇m o
corresponding to the Sauter mean diameter (d32) calculated from

nk {~
T 2
the Feret lengths from the MicroTrac particle analyses. The exit times
out of the top of the reactor were used to create biomass particle
residence time distribution (RTD) curves, which were then averaged (9)
to generate mean transit times. ∂ρ
To determine steady-state particle residence times, it was necessary + ∇(ρμ) = 0
to continue MFIX computations for ∼8 s beyond the initial startup to ∂t (10)
reach a condition where the statistics were stationary. Figure 2 shows The above differential balances can be restated as
∂T
(ρC P)eff = keff ∇(∇T ) − ρC P u∇T + Q
∂t (11)
∂ci
εp + ∇(− Di∇ci) + u∇ci = ∑ Ri (12)
∂t

ρ jijj ∂u u zy
jj + (u∇) zzzz = −∇p
j εp z
k {
l |
εp ∂t
o
o 1 ji o
+ ∇m j zyzo
}
o j
o εp k zo
o {o
o
2
n ~
T
μ(∇ u + (∇ u ) ) − μ(∇ u )I
3
ij y
ρ∇u zzz
− jjjjκij−1μ + zu
j εp2 zz
k { (13)
∂εpρ
+ ∇(ρ u) = ∑ Ri (14)
∂t
where ρ is the density, CP is the heat capacity at a constant pressure, T
is the temperature, k is the thermal conductivity, u is the velocity
vector, Q is the heat generation term, ci is the concentration of species
i, Di is the diffusion coefficient for species i, ∑Ri is the sum reaction
rates for species i, εp is the porosity, μ is the viscosity, I is the identity
matrix, κij is the permeability tensor, and the superscript T in eqs 9
and 13 denotes the transpose operator. The mass source term ∑Ri
represents the reactions from solid wood to vapor products
Figure 2. Reactor hydrodynamic simulation: cross-sectional image of throughout the porous medium. Radiation at the particle surface
char flowing through the FBR with sand. followed the Stefan−Boltzmann equation, with surface emissivity of
0.9 and ambient temperature of 500 °C. Heats of reaction (Table 2)
were applied through the volume of the particle. At the particle
surface, the boundary condition is simply equal flux, such that heat,
a cross section of the 3D reactor simulation for char flowing through mass, and species flux is the same on both sides. A fixed pressure
the fluidized bed with sand. On the basis of this result, total outlet boundary condition follows the equation:
simulation times of 20 s were used to generate statistics for which 2
transition effects could be eliminated.34 − p + μ(∇u + (∇u)T ) − μ(∇u) = − p0
3 (15)
3.2. Particle-Scale Pyrolysis Simulations. Our particle-scale
biomass pyrolysis simulations accounted for previously proposed The intraparticle generation and reaction of pyrolysis vapors were
reaction kinetics coupled to transient, intraparticle differential energy, simulated with finite element integration of the differential balances
species, and momentum balances for the particle characteristics assuming a relative gas flow traveling over each particle surrounded by
described above. Darcy’s law was used to describe the anisotropic a control volume, as illustrated in Figure 3. As explained below, the
fluid transport within porous biomass particles, in which porosity is kinetics used in this study included both “primary” reactions for the
represented as a continuum. The resulting differential balances are initial release of vapors from the solid phase and “secondary” gas-
summarized by phase reactions of the released vapors to create the final products. To
account for secondary vapor-phase reactions outside the particles with
∂T minimal computational effort, the released vapors were assumed to
ρC P = ∇(k∇T ) − ρC P u∇T + Q
∂t (7) have an average total residence time in the reactor based on the time
required for the fluidizing gas to transit from the middle of the bed
∂ci through the entire freeboard. This was determined to be about 1.4 s
+ ∇(− Di∇ci) + u∇ci = R i
∂t (8) from the fluidizing gas conditions used for the experiments of interest.

Table 2. First-Order Arrhenius Reaction Rate Parameters for Pyrolysis Reactions

1 (bio-oil) 2 (gas) 3 (char) 4 (secondary char) 5 (secondary gas) moisture desorption


Ai (s−1) 1.08 × 1010 11 4.38 × 109 11 3.75 × 106 11 1.0 × 105 52 4.28 × 106 53 5.13 × 106 54
Ei (kJ/mol) 14811 152.711 111.711 10853 10853 87.953
ΔH (kJ/kg) 25555 −2055 −2055 −4256 −4256 2700

10686 DOI: 10.1021/acs.energyfuels.8b02309


Energy Fuels 2018, 32, 10683−10694
Energy & Fuels Article

Figure 3. Illustration of the particle-scale modeling approach to account for both primary and secondary pyrolysis reactions.

Table 3. Parameters Used in the Single-Particle Simulations of Pine Pyrolysis


parameter description value reference
Cpwood heat capacity of the wood cell wall 0.1031 + 0.003867T kJ kg−1 K−1 59
Cpchar heat capacity of the char wall 1003.2 + 2.09(T − 273.15) J kg−1 K−1 58
Dbio‑oil diffusion coefficient of naphthalene at 500 °C 0.3 cm2 s−1 60
DN2 diffusion coefficient of nitrogen at 500 °C 1 cm2 s−1 60
ε0 porosity of wood 0.276
ksolid thermal conductivity of the cell wall 0.41 W m−1 K−1 57
kgas thermal conductivity of gas (steam at 500 °C) 0.067 W m−1 K−1
kH2O,L thermal conductivity of liquid water 0.61 W m−1 K−1
kH2O,V thermal conductivity of water vapor at 500 °C and 1.44 atm 0.067 W m−1 K−1
Kchar,axial axial permeability of char (5× wood) 1.17 × 10−10 m2 61
Kchar,radial radial permeability of char (5× wood) 6.67 × 10−15 m2 61
Kwood,axial axial permeability of wood, scaled by the porosity factor (εwood/0.65) 4.69 × 10−12 m2 62
Kwood,radial radial permeability of wood (Kwood,axial/17600) 2.67 × 10−16 m2 62
MC0 initial moisture content 1%
P0 average pressure above the reactor sand bed 109 kPa
Ti initial temperature of the particle at t = 0 298 K
Tr temperature of the reactor and carrier gas 773 K
ρwood density of the wood/char cell wall 1500 kg m−3

Integration of differential biomass particle balances is complicated The reaction rates used to track particle mass loss and complete the
by the change in the properties of each particle as it transits through differential particle balances used the Arrhenius kinetics developed by
the reactor and converts from a raw state to char, as illustrated in Di Blasi,11,52 which track the conversion of pine wood into char and
Figure 1. The limiting average rise velocities and associated residence lumped product classes of light gases and condensable vapors (often
times for particles of raw biomass and fully devolatilized char in the referred to as “tar” or “bio-oil”). These kinetics also account for
depicted computational control volume (denoted U) were deter- secondary reactions, in which condensable vapors crack to light gases
mined from the reactor-scale model for each particle size of wood and and recondense to form additional char.53 Moisture desorption is
char, as described above. As wood is converted to char, the effective represented as an Arrhenius reaction for bound water.54 The
rise velocity increases as the particle density decreases. Thus, for Arrhenius reaction rate parameters used in the present study are
particle-scale simulations, Ui can be effectively estimated by scaling provided in Table 2.
between the limiting rise velocities for each particle size i based on The initial water content in all of the simulations was assumed to
conversion (such that the smallest wood size was converted to the be 1% based on compositional analysis of the feedstock. The porosity,
smallest char size, etc.), as summarized by permeability, and density were calculated at each time step and varied
with the conversion of wood into condensable vapors (bio-oil), light
Ui , t − (1 − α)Ui ,wood + αUi ,char (16) gas, and char. Other properties were held constant, including gas
viscosity, thermal conductivity of solid, and specific heat capacity. The
ρwood viscosity of the gas and any vapor was assumed to be 1 × 10−5 Pa s,
α=1− the viscosity of H2 at 300 °C. Thermal conductivity of the wood cell
ρwood (17)
0 wall (also for char in this study) is 0.41 W m−1 K−1.57 The heat
capacity of wood was taken as [0.1031 + 0.003867T (K−1)] kJ kg−1
where ρwood is the density of partially converted biomass and ρwood0 is
K−1.58
the initial density of biomass before it reacts. The value for gas permeability in non-pelletized wood was 11.2
The varying average particle rise velocities can be related to the darcy units in the longitudinal direction and 4.6 × 10−4 darcy units in
average particle residence times in the reactor (tR) by numerically the radial/tangential direction.62 Permeability scales with a pore size
integrating the particle velocity with respect to time until the particle to the power of 2.63,64 For char, the permeability was taken as that for
moves completely through the reactor to reach the exit at the top (Lt). wood multiplied by a factor of 5.61 The permeability was reduced by a
tR factor of 4 for the pelletized wood based on pore size analysis. The
Lt = ∫0 Ui , t dt
(18)
permeability at a given time at a point in the particle is determined by
K wood, t ρwood Kchar, t (1 − ρwood )
Thus, when the limiting particle rise velocities for raw biomass and Kt = +
fully devolatilized char are available (e.g., from reactor CFD ρwood ρwood (19)
0 0
simulations), it is possible to use the above relationship to relate
the average residence time of each particle to its average conversion at where K is permeability, ρwood is the density of wood evolving with
the reactor exit. conversion, and ρwood0 is the initial density of wood in the particle.

10687 DOI: 10.1021/acs.energyfuels.8b02309


Energy Fuels 2018, 32, 10683−10694
Energy & Fuels Article

Figure 4. (A) Biomass and char particle size distributions from the MicroTrac measurements. For convenience in the simulations described below,
these distributions were split into five bins spanning the sizes and shapes present in the feedstock. (B) Mean width and thickness of particles for
their respective particle ferret lengths used to construct single-particle representations.

Figure 5. X-ray computed tomographic data comparing the structure of (left) milled and (right) pelletized/crushed pine feedstock particles.

Additional key parameters used in the single-particle model are shown ε0(ρwood + ρchar )
in Table 3. ε=1−
ρwood (21)
The bulk diffusivity (Di) of bio-oil was taken to be the diffusivity of 0

naphthalene in nitrogen at 500 °C, and Di of gases and nitrogen was where ε is the porosity and ρchar is the density of char. The density of
taken to be the diffusivity of nitrogen at the same temperature.60 the solid porous matrix was calculated by the following equation:
Inside the particle, the effective diffusivity Deff,i is based on the bulk ρsolid = ρwood + ρchar + ρH O
0 2 L (22)
diffusivity and porosity65
The density of vapor/gas was calculated by the following equation
ε (Dalton’s law):
Deff, i = Di
τ (20) p
ρgas = (Cgas + C tar + C N2 + C H2OV )
RT (23)
where ε is the porosity and τ is the tortuosity calculated by the
Millington and Quirk model66 to be ε−1/3. The initial porosities of where R is the universal gas constant, p is the pressure, T is the
non-densified pine and poplar are 0.65 and 0.80, respectively, based temperature, and Ci is the molar concentration of each species.
on the cell wall density of 1500 kg/m3 and reported bulk wood Reaction-coordinate-dependent shrinking was accounted for in the
reactor-scale model but not in the single-particle model because it has
densities.58 On the basis of the X-ray tomography data, the initial
been previously shown to have little impact at the particle scale.38
porosity of our densified pine is 0.270. The porosity at a given time at 3.3. Canonical Ensemble Yield Calculations. As discussed
a point in the particle follows eq 21. The impact of the moisture above, the as-fed biomass particles had a relatively wide range of sizes.
content on the porosity was ignored because the rate of pyrolysis To account for this, five representative particle sizes were identified to
reactions is negligible below ∼350 °C span the range of experimental sizes for reactor- and particle-scale

10688 DOI: 10.1021/acs.energyfuels.8b02309


Energy Fuels 2018, 32, 10683−10694
Energy & Fuels Article

simulations. Results for each size were then used to compute volume-
weighted averages for the expected bio-oil, gas, char, and unconverted
biomass yields predicted by the models that could be compared to
experimental measurements. This is summarized below, where y is the
yield of bio-oil, char, gas, secondary char, or biomass (wood), V is the
volume fraction, and i is the bin number.
i = 1:5
y= ∑ Vy
ii
bins (24)

4. RESULTS AND DISCUSSION


4.1. Experimental Feedstock Analysis. The as-fed
biomass particles have a relatively wide range of sizes, even
for the milled pellets, which have been screened to 2 mm prior
to feeding, as seen in Figure 4. It also appears that the char
particles exiting the reactor are typically much smaller than the
as-fed wood particles, with the most populous particle size
dropping from 2.3 to 0.4 mm in length. We speculate that the
harsh environment of the sand fluidized bed likely causes
significant attrition and fragmentation of the biomass particles
as they transit the reactor. Figure 6. Linear dependence of thermal conductivity upon specific
The XCT results provided porosity and pore structure as gravity.58
well as a basis for estimating the effective thermal conductivity
of the milled and pelletized biomass particles because the sizes of raw biomass and char particles were used to determine
thermal conductivity of woody biomass is linearly dependent the RTDs and mean reactor transit times for each particle class.
upon its specific gravity, which is dictated by the presence of Table 4 and Figure 7 summarize the results for both raw
void space resulting from the cellular tissue structure of
biomass.58 This dependence is illustrated in, for example, Table 4. Wood and Char Particle Residence Times for the
woody materials in Figure 5. Thus, we were able to estimate Five Representative Particle Sizesa
the original particle thermal conductivity from the XCT images
using the interior void fraction (denoted f v) that was revealed Feret length (mm) particle residence time (s)
by dividing the filled tomographic volume of the particle wood 0.76 2.67
interior by the total volume encapsulated by the exterior 1.52 -
particle boundary. Because the specific gravity of the cell wall is 2.27 -
largely constant at ∼1.5 among wood species,67 the total 3.03 -
density of the particle was estimated as ρ = (1 − f v)1.5 (g/ 4 -
cm3). The thermal conductivity was then estimated using the char 0.11 2.45
linear expression shown below the fitted line in Figure 6 by 0.44 3.31
generating a weighted average of the void space domain (kgas) 0.78 3.70
and the solid domain (ksolid) in the particles. 1.08 4.22
Another useful result from the XCT analyses is that the feed 1.44 4.69
a
preparation steps used in the experiments increased the Note that for all but the smallest wood particle size, the particle is
effective feed particle density from ∼540 kg/m3 for the milled predicted to not leave the reactor, as represented by “-”.
pine to ∼1096 kg/m3 for pelletized pine. This corresponds to
changes in the effective thermal conductivity from 0.12 to 0.23 biomass (pine wood) and char particles at various sizes based
W m−1 K−1 for the milled and pelletized feedstock, on the particle size distribution from the particle analyses
respectively. While it has been speculated that densification described above.
of biomass during pelletization can impact the particle As expected, the MFIX simulation predicted that larger and
microstructure and internal particle transport properties, this more dense particles had increased residence times in the
is the first time to our knowledge that the phenomenon has reactor. In cases where the as-fed biomass particles were
been experimentally verified. sufficiently large, it was not even possible to determine finite
Milling also appears to have an impact on the distribution of reactor residence times within the limits of the current MFIX
the biomass particle aspect ratio. Specifically, the milling simulations. This should not be surprising, because the physics
process used in the NREL experiments produces relatively of particle elutriation require that there is a critical size limit for
high-aspect particles that preserve the inherent directional biomass particles such that they are never able to elutriate from
microstructure. In contrast, XCT images of the pelletized feed the reactor unless the gas flow is increased. Conversely, the
particles appear to suggest more of a random agglomeration of smallest particles are almost able to attain the velocity of the
smaller particles, which has less net directional character. gas (near zero slip velocity), resulting in very short residence
4.2. Reactor-Scale Modeling. CFD modeling of the times nearly equivalent to the gas phase. The presence of such
2FBR with MFIX was used to estimate the residence times of critical limits illustrates the kind of strong nonlinear relation-
biomass and char particles to provide heating and reaction ships involved in selecting design and operating parameters,
time scales that could be applied to particle-scale pyrolysis such as particle size and gas flow, that need to be considered in
simulations. Specifically, the simulated behavior of selected optimizing pyrolysis FBRs.
10689 DOI: 10.1021/acs.energyfuels.8b02309
Energy Fuels 2018, 32, 10683−10694
Energy & Fuels Article

estimated for these same particle sizes, as summarized in Table


5. Notably, the 0.76 mm particle, which has a residence time of
2.6 s, only achieves 87% conversion because its small size
makes it pass very quickly through the reactor before the
primary reactions can be completed. This also results in a
reduced predicted bio-oil yield of 49.6% by weight. On the
other hand, the 3.79 mm particle has a 9.3 s residence time,
allowing the primary reactions to be about 97% completed and
a higher predicted bio-oil yield of 55−56% by weight. Again,
these differences illustrate the importance of the particle size
on the pyrolysis performance. Of course, size effects should be
weighted by the relative abundance of each size. For example,
although the conversion of the 0.76 mm particles is relatively
high, these particles only account for about 1.1% of the
feedstock on a weight basis, and thus, they do not dramatically
affect the overall biomass conversion.
Figure 7. Superficial gas−particle slip velocity of gas around relevant The predicted exit product yields from our simulations for
particle sizes (dashed lines) and average limiting particle rise velocity the 2FBR appear to agree well with the available experimental
through the FBR (solid lines). values (within the estimated error limits), as also shown in
Table 5. This suggests to us that the basic computational
4.3. Pyrolysis Yields from Coupled Reactor- and modeling approach that we have proposed for combining
Single-Particle-Scale Models. The particle-scale models computational simulations of reactor- and particle-scale
captured intraparticle heat conduction, mass diffusion, mass processes is sufficiently detailed to capture the major effects
convection, pressure, density, chemical reactions, and solid− associated with feed particle variations, without the addition of
gas phase change. Particle-scale simulations were run for each arbitrarily fit parameters.
of the five representative pine particles over a range of reaction 4.4. Additional Particle-Scale Simulations. To evaluate
times, to produce the conversion snapshots after 4 s, illustrated the practical value of the particle-scale model implemented
in Figure 8. One notable observation revealed by these here for process development, we also used it to explore the
simulations is that rather high velocities (up to 0.25 m/s) were predicted impact of changes in specific pyrolysis parameters.
calculated for the pyrolysis vapors that were ejected from the Initially, we did this by considering the impact of changes in
ends of the porous pine particle. The average particle the reactor temperature and particle residence time. Assuming
temperature and yields predicted by the particle-scale model a single feed particle size of 2.27 mm and a constant gas
as a function of the exposure time in the reactor for different residence time of 1.4 s, the trends in the bio-oil yield and the
size particles are summarized in Figure 9. The temperature relationship between the reactor temperature and particle
results illustrate that initial heating is very rapid for the smallest residence time illustrated in Figure 10 were generated. These
particles (0.76 mm long), and they reach 95% of their final results imply that there is an optimum combination of
temperature after only 0.7 s. However, the largest particles (3.8 temperature and particle residence time in pyrolysis reactors
mm long) take 4.9 s to achieve the same temperature, revealing that should produce the maximum bio-oil yield. In the specific
how dramatic the effect of feed particle size is on the heating example depicted, this optimal bio-oil yield is predicted to
rate. This effect is due to a combination of the lower thermal approach 69.5% by weight at a reactor temperature of 430 °C
mass and higher surface area/volume ratio for the smaller (703 K). Such a yield is considerably higher than that achieved
particles. to date in experiments with this FBR, indicating that significant
On the basis of the mean residence time predictions from potential remains for additional process improvement.
the reactor-scale simulations combined with the residence As another example, we considered the joint impact of the
time−conversion relationship represented by eq 18 and the gas residence time and temperature changes on the bio-oil
assumptions concerning secondary vapor-phase reactions yield, shown in Figure 11. The impact of the gas residence time
described above, the product yields at the reactor exit were is enhanced as the temperature increases, particularly above

Figure 8. Predicted wood conversion and exit gas velocities for five different size particles after 4 s.

10690 DOI: 10.1021/acs.energyfuels.8b02309


Energy Fuels 2018, 32, 10683−10694
Energy & Fuels Article

Figure 9. Yields (wt/wt from dry wood) of pyrolysis products from five selected particle sizes versus reaction time.

Table 5. Summary of Simulation Results Compared to Experimental Resultsa


bins volume-weighted average experimental yields
Feret length of wood (mm) 0.76 1.52 2.27 3.03 3.79
volume percent 1.1 34.3 50.7 12.0 1.9
Feret length of char (mm) 0.11 0.44 0.78 1.08 1.44
effective particle residence time (s) 2.6 5.4 6.6 7.9 9.3
gas residence time (s) 1.4 1.4 1.4 1.4 1.4
bio-oil yield 49.6 55.9 55.6 55.3 55.2 55.6 55.6 ± 1.3
char yield 7.85 9.3 9.77 10.1 10.5 9.2 11.3 ± 0.5
gas + water vapor yield 29.5 33.2 33.0 32.7 32.7 33.0 30.9 ± 1.7
wood conversion 87.2 97.6 97.3 97.0 97.1 97.2
a
Yields in wt/wt % from dry wood. Experimental error represents population standard deviation.

Figure 10. (Left) Effect of the particle residence time and temperature on the yield of bio-oil with varying particle residence times and (right)
optimal temperature versus particle residence time with a fixed 1.4 s gas residence time for the 2.27 mm particle.

560 °C (840 K), where the rate of secondary reactions is very decomposed to light gases in less than 0.1 s. Above this
high and over half of the initially generated bio-oil is temperature and with gas residence times greater than 2 s,
10691 DOI: 10.1021/acs.energyfuels.8b02309
Energy Fuels 2018, 32, 10683−10694
Energy & Fuels Article

Figure 11. (Left) Bio-oil yield and (right) gas yield with the temperature and gas residence time with a fixed particle residence time of 6.6 s for a
2.27 mm particle.

nearly all of the bio-oil is predicted to be converted into gas. simultaneously (e.g., such as CFD simulations that directly
Interestingly, these results also suggest that there is a regime couple hydrodynamics with chemical reactions37,38). For other
with high bio-oil yields (greater than 50 wt/wt %) that is less more complex process contexts, it may be necessary to account
sensitive to the gas residence time between 400 and 450 °C more fully for multidirectional coupling among the scales
(673−723 K). (rather than the one-way coupling implemented in this work),
Clearly, these above examples illustrate the usefulness of but the potential advantages of making the information transfer
such particle-scale models for planning experiments, designing process more selective between separate, independently run
reactors, and assessing the technoeconomic potential of models would appear to be considerable.
candidate processes. While the design and operating
limitations of the reference reactor considered in this study
may prohibit the realization of the predicted optimal
■ AUTHOR INFORMATION
Corresponding Author
conditions, the computational simulations make it possible to *Telephone: 1-303-384-7691. E-mail: peter.ciesielski@nrel.
recognize important opportunities for future research. gov.
ORCID
5. CONCLUSION Gavin M. Wiggins: 0000-0002-4737-6596
The results reported here demonstrate that combined Peter N. Ciesielski: 0000-0003-3360-9210
implementation of separate reactor- and particle-scale models Funding
can lead to new insights and opportunities for understanding This work was authored by the Alliance for Sustainable Energy,
and improving biomass fast pyrolysis in fluidized beds. At the LLC, the manager and operator of the National Renewable
reactor scale, it is feasible to simulate the biomass and char Energy Laboratory for the U.S. Department of Energy (DOE)
particle residence times with sufficient accuracy using two-fluid under Contract DE-AC36-08GO28308. Funding was provided
CFD at a reasonable cost in time and computing resources. At by the U.S. Department of Energy Office of Energy Efficiency
the particle scale, it is feasible to integrate the differential and Renewable Energy Bioenergy Technologies Office. The
intraparticle heat, mass, and momentum balances with finite views expressed in the article do not necessarily represent the
element methods (FEMs) for selected classes of particles while views of the DOE or the U.S. Government. The U.S.
still including realistic assumptions about feedstock particle Government retains and the publisher, by accepting the article
size variations and microstructure. Taken together, the for publication, acknowledges that the U.S. Government
combined information generated by independent simulations retains a nonexclusive, paid-up, irrevocable, worldwide license
with each of these models can be used to accurately predict the to publish or reproduce the published form of this work or
experimentally observed product yields from the NREL 2 in. allow others to do so for U.S. Government purposes.
fluidized bed pyrolyzer within measurement accuracy. As Notes
additional kinetic details for the pyrolysis reactions of different
The authors declare no competing financial interest.


types of biomass become available, we expect that additional
experimental verifications for using this modeling approach will ACKNOWLEDGMENTS
be made.
For future work, we recommend continued utilization of this The authors acknowledge Steve Deutch and Kellene Orton for
type of multimodel simulation strategy for simulating multi- information and discussions about the FBR experimental
setup.


scale biomass conversion processes. Combining information
generated by separate models that approximate the dominant
physics at different time and length scales would appear to NOMENCLATURE
offer significant advantages in computational efficiency over CFD = computational fluid dynamics
models that attempt to address all of the scales in detail FBR = fluidized bed reactor
10692 DOI: 10.1021/acs.energyfuels.8b02309
Energy Fuels 2018, 32, 10683−10694
Energy & Fuels Article

FEM = finite element method (18) Ciesielski, P. N.; Crowley, M.; Nimlos, M.; Sanders, A.;
MFIX = multiphase flow with interphase exchanges, a Wiggins, G.; Robichaud, D.; Donohoe, B.; Foust, T. Biomass Particle
reactor modeling software package Models with Realistic Morphology and Resolved Microstructure for
NREL = National Renewable Energy Laboratory Simulations of Intraparticle Transport Phenomena. Energy Fuels 2015,
XCT = X-ray computed tomography 29 (1), 242−254.


(19) Okekunle, P. O.; Watanabe, H.; Pattanotai, T.; Okazaki, K.
Effect of Biomass Size and Aspect Ratio on Intra-Particle Tar
REFERENCES Decomposition during Wood Cylinder Pyrolysis. J. Therm. Sci.
(1) Hu, W.; Dang, Q.; Rover, M.; Brown, R. C.; Wright, M. M. Technol. 2012, 7 (1), 1−15.
Comparative techno-economic analysis of advanced biofuels, (20) Mathews, J. P.; Campbell, Q. P.; Xu, H.; Halleck, P. A review of
biochemicals, and hydrocarbon chemicals via the fast pyrolysis the application of X-ray computed tomography to the study of coal.
platform. Biofuels 2016, 7 (1), 57−67. Fuel 2017, 209, 10−24.
(2) Ciesielski, P. N.; Pecha, M. B.; Bharadwaj, V. S.; Mukarakate, C.; (21) Hounsfield, G. N. Computerized transverse axial scanning
Leong, G. J.; Kappes, B.; Crowley, M. F.; Kim, S.; Foust, T. D.; (tomography): Part 1. Description of system. Br. J. Radiol. 1973, 46
Nimlos, M. R. Advancing catalytic fast pyrolysis through integrated (552), 1016−1022.
multiscale modeling and experimentation: Challenges, progress, and (22) Flannery, B. P.; Deckman, H. W.; Roberge, W. G.; D’Amico, K.
perspectives. Wiley Interdiscip. Rev.: Energy Environ. 2018, 7 (4), e297. L. Three-dimensional X-ray microtomography. Science 1987, 237
(3) Mettler, M. S.; Vlachos, D. G.; Dauenhauer, P. J. Top Ten (4821), 1439−1444.
Fundamental Challenges of Biomass Pyrolysis for Biofuels. Energy (23) Flannery, B. P.; Roberge, W. Observational strategies for three-
Environ. Sci. 2012, 5 (7), 7797−7809. dimensional synchrotron microtomography. J. Appl. Phys. 1987, 62
(4) Garcia-Nunez, J. A.; Pelaez-Samaniego, M. R.; Garcia-Perez, M. (12), 4668−4674.
E.; Fonts, I.; Abrego, J.; Westerhof, R. J. M.; Garcia-Perez, M. (24) Hyväluoma, J.; Kulju, S.; Hannula, M.; Wikberg, H.; Källi, A.;
Historical Developments of Pyrolysis Reactors: A Review. Energy Rasa, K. Quantitative characterization of pore structure of several
Fuels 2017, 31 (6), 5751−5775. biochars with 3D imaging. Environ. Sci. Pollut. Res. 2018, 25 (26),
(5) Kersten, S. R. A.; Wang, X.; Prins, W.; van Swaaij, W. P. M. 25648−25658.
Biomass Pyrolysis in a Fluidized Bed Reactor. Part 1: Literature (25) Jones, K.; Ramakrishnan, G.; Uchimiya, M.; Orlov, A. New
Review and Model Simulations. Ind. Eng. Chem. Res. 2005, 44 (23), Applications of X-ray Tomography in Pyrolysis of Biomass: Biochar
8773−8785. Imaging. Energy Fuels 2015, 29 (3), 1628−1634.
(6) Wang, X.; Kersten, S. R. A.; Prins, W.; van Swaaij, W. P. M. (26) Watanabe, H. X-ray Computed Tomography Visualization of
Biomass Pyrolysis in a Fluidized Bed Reactor. Part 2: Experimental the Woody Char Intraparticle Pore Structure and Its Role on
Validation of Model Results. Ind. Eng. Chem. Res. 2005, 44 (23), Anisotropic Evolution during Char Gasification. Energy Fuels 2018, 32
8786−8795. (4), 4248−4254.
(7) Blanco, A.; Chejne, F. Modeling and simulation of biomass fast (27) Di Blasi, C. Heat, momentum and mass transport through a
pyrolysis in a fluidized bed reactor. J. Anal. Appl. Pyrolysis 2016, 118, shrinking biomass particle exposed to thermal radiation. Chem. Eng.
105−114. Sci. 1996, 51 (7), 1121−1132.
(8) Di Blasi, C. Modeling Chemical and Physical Processes of Wood (28) Bellais, M. Modelling of the pyrolysis of large wood particles.
and Biomass Pyrolysis. Prog. Energy Combust. Sci. 2008, 34 (1), 47− Ph.D. Thesis, KTH Royal Institute of Technology, Stockholm,
90. Sweden, 2007.
(9) Debiagi, P. E. A.; Gentile, G.; Pelucchi, M.; Frassoldati, A.; (29) Paulsen, A. D.; Hough, B. R.; Williams, C. L.; Teixeira, A. R.;
Cuoci, A.; Faravelli, T.; Ranzi, E. Detailed kinetic mechanism of gas- Schwartz, D. T.; Pfaendtner, J.; Dauenhauer, P. J. Fast Pyrolysis of
phase reactions of volatiles released from biomass pyrolysis. Biomass Wood for Biofuels: Spatiotemporally Resolved Diffuse Reflectance In
Bioenergy 2016, 93, 60−71. situ Spectroscopy of Particles. ChemSusChem 2014, 7 (3), 765−776.
(10) Diebold, J. P. A unified, global model for the pyrolysis of (30) Oh, M. S.; Peters, W. A.; Howard, J. B. An experimental and
cellulose. Biomass Bioenergy 1994, 7 (1), 75−85. modeling study of softening coal pyrolysis. AIChE J. 1989, 35 (5),
(11) Di Blasi, C.; Branca, C. Kinetics of primary product formation 775−792.
from wood pyrolysis. Ind. Eng. Chem. Res. 2001, 40 (23), 5547−5556. (31) Sezen, Y. Internal mass transfer considerations during the
(12) Corbetta, M.; Frassoldati, A.; Bennadji, H.; Smith, K.; pyrolysis of an isolated spherical softening coal particle. Int. J. Heat
Serapiglia, M. J.; Gauthier, G.; Melkior, T.; Ranzi, E.; Fisher, E. M. Mass Transfer 1989, 32 (10), 1992−1987.
Pyrolysis of Centimeter-Scale Woody Biomass Particles: Kinetic (32) Teixeira, A. R.; Mooney, K. G.; Kruger, J. S.; Williams, C. L.;
Modeling and Experimental Validation. Energy Fuels 2014, 28 (6), Suszynski, W. J.; Schmidt, L. D.; Schmidt, D. P.; Dauenhauer, P. J.
3884−3898. Aerosol generation by reactive boiling ejection of molten cellulose.
(13) Ranzi, E.; Debiagi, P. E. A.; Frassoldati, A. Mathematical Energy Environ. Sci. 2011, 4 (10), 4306−4321.
modeling of fast biomass pyrolysis and bio-oil formation. Note I: (33) Ghazaryan, L. Aerosol dynamics in porous media. Ph.D. Thesis,
Kinetic mechanism of biomass pyrolysis. ACS Sustainable Chem. Eng. University of Twente, Twente, Netherlands, 2014.
2017, 5 (4), 2867−2881. (34) Ramirez, E.; Finney, C. E. A.; Pannala, S.; Daw, C. S.; Halow, J.;
(14) Ranzi, E.; Debiagi, P. E. A.; Frassoldati, A. Mathematical Xiong, Q. Computational study of the bubbling-to-slugging transition
Modeling of Fast Biomass Pyrolysis and Bio-Oil Formation. Note II: in a laboratory-scale fluidized bed. Chem. Eng. J. 2017, 308, 544−556.
Secondary Gas-Phase Reactions and Bio-Oil Formation. ACS (35) Xiong, Q.; Yang, Y.; Xu, F.; Pan, Y.; Zhang, J.; Hong, K.;
Sustainable Chem. Eng. 2017, 5 (4), 2882−2896. Lorenzini, G.; Wang, S. Overview of Computational Fluid Dynamics
(15) Ciesielski, P. N.; Wiggins, G. M.; Jakes, J. E.; Daw, C. S. Simulation of Reactor-Scale Biomass Pyrolysis. ACS Sustainable Chem.
Simulating Biomass Fast Pyrolysis at the Single Particle Scale. Fast Eng. 2017, 5 (4), 2783−2798.
Pyrolysis of Biomass 2017, 231−253. (36) Papadikis, K.; Gu, S.; Bridgwater, A. V. Computational
(16) Shen, J.; Wang, X.-S.; Garcia-Perez, M.; Mourant, D.; Rhodes, modelling of the impact of particle size to the heat transfer coefficient
M. J.; Li, C.-Z. Effects of particle size on the fast pyrolysis of oil mallee between biomass particles and a fluidised bed. Fuel Process. Technol.
woody biomass. Fuel 2009, 88 (10), 1810−1817. 2010, 91 (1), 68−79.
(17) Maa, P. S.; Bailie, R. C. Influence of Particle Sizes and (37) Papadikis, K.; Gu, S.; Bridgwater, A. V. CFD modelling of the
Environmental Conditions on High Temperature Pyrolysis of fast pyrolysis of biomass in fluidised bed reactors. Part B: Heat,
Cellulosic MaterialI (Theoretical). Combust. Sci. Technol. 1973, 7, momentum and mass transport in bubbling fluidised beds. Chem. Eng.
257−269. Sci. 2009, 64 (5), 1036−1045.

10693 DOI: 10.1021/acs.energyfuels.8b02309


Energy Fuels 2018, 32, 10683−10694
Energy & Fuels Article

(38) Papadikis, K.; Gu, S.; Bridgwater, A. V. CFD modelling of the (59) Juřena, T. Numerical modelling of grate combustion. Ph.D.
fast pyrolysis of biomass in fluidised bed reactors: Modelling the Thesis, Brno University of Technology, Brno, Czech Republic, 2012.
impact of biomass shrinkage. Chem. Eng. J. 2009, 149 (1), 417−427. (60) Chae, K. Mass diffusion and chemical kinetic data for jet fuel
(39) Xiong, Q.; Kong, S.-C.; Passalacqua, A. Development of a surrogates. Ph.D. Dissertation, University of Michigan, Ann Arbor,
generalized numerical framework for simulating biomass fast pyrolysis MI, 2010.
in fluidized-bed reactors. Chem. Eng. Sci. 2013, 99 (0), 305−313. (61) Di Blasi, C. Modelling the fast pyrolysis of cellulosic particles in
(40) Taghipour, F.; Ellis, N.; Wong, C. Experimental and fluid-bed reactors. Chem. Eng. Sci. 2000, 55 (24), 5999−6013.
computational study of gas−solid fluidized bed hydrodynamics. (62) Comstock, G. L. Directional permeability of softwoods. Wood
Chem. Eng. Sci. 2005, 60 (24), 6857−6867. Fiber Sci. 1970, 1 (4), 283−289.
(41) Xue, Q.; Fox, R. O. Computational modeling of biomass (63) Bliek, A.; Van Poelje, W.; Van Swaaij, W.; Van Beckum, F.
thermochemical conversion in fluidized beds: Particle density Effects of intraparticle heat and mass transfer during devolatilization
variation and size distribution. Ind. Eng. Chem. Res. 2015, 54 (16), of a single coal particle. AIChE J. 1985, 31 (10), 1666−1681.
4084−4094. (64) Morell, J. I.; Amundson, N. R.; Park, S.-K. Dynamics of a single
(42) Anca-Couce, A. Reaction mechanisms and multi-scale particle during char gasification. Chem. Eng. Sci. 1990, 45 (2), 387−
modelling of lignocellulosic biomass pyrolysis. Prog. Energy Combust. 401.
Sci. 2016, 53, 41−79. (65) Mason, E. A.; Malinauskas, A. Gas Transport in Porous Media:
(43) Sharma, A.; Pareek, V.; Zhang, D. Biomass pyrolysisA review The Dusty-Gas Model; Elsevier Science, Ltd.: Amsterdam, Nether-
of modelling, process parameters and catalytic studies. Renewable lands, 1983; Vol. 17.
Sustainable Energy Rev. 2015, 50, 1081−1096. (66) Millington, R.; Quirk, J. Permeability of porous solids. Trans.
(44) Xiong, Q.; Xu, F.; Pan, Y.; Yang, Y.; Gao, Z.; Shu, S.; Hong, K.; Faraday Soc. 1961, 57, 1200−1207.
Bertrand, F.; Chaouki, J. Major trends and roadblocks in CFD-aided (67) Kellogg, R. M.; Wangaard, F. F. Variation in the cell-wall
process intensification of biomass pyrolysis. Chem. Eng. Process. 2018, density of wood. Wood Fiber Sci. 1969, 1 (3), 180−204.
127, 206−212.
(45) Wurzenberger, J. C.; Wallner, S.; Raupenstrauch, H.; Khinast, J.
G. Thermal conversion of biomass: Comprehensive reactor and
particle modeling. AIChE J. 2002, 48 (10), 2398−2411.
(46) Howe, D.; Westover, T.; Carpenter, D.; Santosa, D.; Emerson,
R.; Deutch, S.; Starace, A.; Kutnyakov, I.; Lukins, C. Field-to-Fuel
Performance Testing of Lignocellulosic Feedstocks: An Integrated
Study of the Fast Pyrolysis−Hydrotreating Pathway. Energy Fuels
2015, 29 (5), 3188−3197.
(47) MicroTrac. PartAn 3-DDynamic Image Analyzer; MicroTrac:
Montgomeryville, PA, 2018; https://www.microtrac.com/partan-3-d-
particle-size-and-shape-analyzer/.
(48) Syamlal, M.; Rogers, W.; O’Brien, T. J. MFIX Documentation
Theory Guide; Morgantown Energy Technology Center, U.S.
Department of Energy: Morgantown, WV, 1993.
(49) Syamlal, M.; O’Brien, T. A Generalized Drag Correlation for
Multiparticle Systems; Office of Fossil Energy, U.S. Department of
Energy: Washington, D.C., 1987.
(50) Schaeffer, D. G. Instability in the evolution equations
describing incompressible granular flow. Journal of Differential
Equations 1987, 66 (1), 19−50.
(51) Pannala, S.; Daw, C. S.; Finney, C. E. A.; Benyahia, S.; Syamlal,
M.; O’Brien, T. J.; Nakagawa, M.; Luding, S. Modeling the
Collisional-Plastic Stress Transition for Bin Discharge of Granular
Material. AIP Conf. Proc. 2009, 1145 (1), 657−660.
(52) Di Blasi, C. Analysis of Convection and Secondary Reaction
Effects Within Porous Solid Fuels Undergoing Pyrolysis. Combust. Sci.
Technol. 1993, 90 (5−6), 315−340.
(53) Liden, A. G.; Berruti, F.; Scott, D. S. A kinetic model for the
production of liquids from the flash pyrolysis of biomass. Chem. Eng.
Commun. 1988, 65 (1), 207−221.
(54) Gezici-Koç, Ö .; Erich, S. J. F.; Huinink, H. P.; van der Ven, L.
G. J.; Adan, O. C. G. Bound and free water distribution in wood
during water uptake and drying as measured by 1D magnetic
resonance imaging. Cellulose 2017, 24 (2), 535−553.
(55) Koufopanos, C. A.; Papayannakos, N.; Maschio, G.; Lucchesi,
A. Modelling of the pyrolysis of biomass particles. Studies on kinetics,
thermal and heat transfer effects. Can. J. Chem. Eng. 1991, 69 (4),
907−915.
(56) Chan, W.-C. R.; Kelbon, M.; Krieger, B. B. Modelling and
experimental verification of physical and chemical processes during
pyrolysis of a large biomass particle. Fuel 1985, 64 (11), 1505−1513.
(57) Hunt, J. F.; Gu, H.; Lebow, P. K. Theoretical thermal
conductivity equation for uniform density wood cells. Wood Fiber Sci.
2008, 40 (2), 167−180.
(58) Ross, R. J. Wood Handbook, Wood as an Engineering Material;
Forest Products Laboratory: Madison, WI, 2010.

10694 DOI: 10.1021/acs.energyfuels.8b02309


Energy Fuels 2018, 32, 10683−10694

Вам также может понравиться