You are on page 1of 18

European Combustion Meeting (ECM2003) Plenary Lecture

Biomass Combustion
J Swithenbank, YB Yang, C Ryu, J Goodfellow,
S Shabangu, N V Russell, F M Lewis*, V N Sharifi
Sheffield University Waste Incineration Centre (SUWIC), Department of Chemical and
Process Engineering, University of Sheffield, Sheffield S1 3JD, UK
* 535 East Mariposa Ave, El Segundo, CA 90245- 3013, USA
ABSTRACT
Sustainable cities require the generation of electrical energy from carbon dioxide neutral biomass crops and
suitable fractions of wastes that cannot be economically reused or recycled. The energy content of these
solid materials can be recovered by burning directly or after processing into refuse-derived fuel (RDF).
Alternatively, the combustion process can be staged by the production of intermediate fuels using either
pyrolysis or gasification. Co-processing of the biomass with coal generally increases plant utilisation and
thus reduces costs.
The design, operation and maintenance of solid fuel combustion, pyrolysis or gasification plants requires
detailed understanding of the processes occurring within a reacting packed bed of solids, or combustion of
the derived liquids or gaseous fuels. Both of the latter can be modelled using well-established computational
fluid dynamic codes (CFD). Previously, there has not been available a validated, comprehensive and
fundamentally based code for mathematically modelling the combustion/pyrolysis/gasification process
within a packed bed of solid particles on a stationary or moving grate. Bearing in mind that pyrolysis and
gasification are sub-sets of combustion we have developed a generalised model of bed combustion. This
code known as FLIC solves iteratively the flow field within a reacting bed of randomly packed particles,
including radiant heat transfer. The equations governing the processes of drying, pyrolysis, de-volatilisation
and char burnout within the particles are evaluated. Since the burning of volatiles and CO in the channels is
mixing limited, flame reactions also occur in the gas phase above the bed. The conditions evaluated at the
surface of the bed are the boundary conditions for conventional CFD modelling of the mixing and reactions
in the secondary combustion zone in the freeboard above the bed and in the gas clean-up system. This
permits the evaluation and minimization of emissions such CO, VOCs, NOx, heavy metals and dioxins. In
fact, dioxins from incinerators now only contribute 3% of the total UK dioxin emissions.
Pyrolysis of biomass/waste can generate a storable char fuel. This is achieved by heating a bed of the
material slowly in a closed container to about 500C from which air is excluded. This decomposes the
organic material to release liquid products that can be condensed then purified and burned to efficiently
generate heat and power. The carbon char remaining is a valuable fuel that can be easily separated whilst it
is still hot from any inert material that was originally present. This storable fuel can be transported and used
when the renewable energy that it contains is required. The char from pyrolysis contains much of the
original carbon and has a high-energy content.
The validation of our reacting bed modelling code (FLIC) has been achieved by measurements in a pot
burner using various biomass materials. Additionally, a small ball instrument that has been specially
developed to contain instruments has complemented these measurements by withstanding temperatures up to
1000oC for well over an hour. This novel device passes through industrial moving grate furnaces with the
fuel and records parameters such as oxygen concentration, vibration and several temperatures onto a
computer memory chip. The ball is recovered from the ash pit and the information is downloaded onto an
Excel spreadsheet for detailed analysis and verification of the FLIC predictions.
Our new gasification concept offers the prospect of achieving the goal of high efficiency power generation
from char, coal or various biomass sources by utilizing ultra superheated steam (USS). The method uses
low-grade steam from sources such as process cooling, waste incineration and local industry, and then
enhances it to a temperature above 1600C. This is achieved by adding oxygen to the steam to form
"artificial air"; gas is then burned in this artificial air to produce an ultra-superheated steam flame. Char,
biomass or powdered coal are injected coaxially into the high temperature steam flame where they react in
about a second to produce a gas that is free of tar and consists largely of CO, H2 and CH4. Significantly, the
high temperature steam provides the enthalpy for the endothermic gasification reactions. This feature
ensures that tar formation is avoided and any ash is cooled below its melting point before it arrives at the
reactor walls. This process provides a route to the future hydrogen fuel based economy.
These innovative technologies form part of an integrated consortium programme towards sustainability.

1
BACKGROUND
Worldwide, there is a progressive change to use a greater proportion of our total energy
consumption as electrical energy and citizens in developed countries now use up to about
1 kW of electricity per person. This corresponds to about 2.4 tonnes of coal per person per
year. At the same time, citizens bring into being up to one tonne of municipal waste each
year having an energy equivalent of 300 kg of coal plus even larger quantities of agricultural
and industrial wastes. Bearing in mind that these wastes are largely biomass, the recovery of
energy from the waste would not add very much to the net atmospheric CO2 level. Thus
recovering energy from waste helps to mitigate the climate change problem. Furthermore,
dedicated agricultural production of biomass for energy generation is increasing in popularity.
Sustainable cities therefore require the generation of electrical energy from biomass including
suitable fractions of waste that cannot be economically reused or recycled. The integration of
these fuels with clean coal technology can also result in better utilization of the generation
plant and thus reduce the cost of the electricity produced.

INTRODUCTION

The combustion of biomass can be accomplished either on a grate, in a fluidised bed, or as


entrained particles. This presentation addresses the phenomena involved in their combustion
on a static or moving grate. The four
principle stages involved are: drying,
pyrolysis/devolatilisation, oxidation Bed Combustion Zones
of the volatile material and char
combustion, finally leaving any
Drying
residual ash. These processes can
take place in separate stages as in Pyrolysis
Oxidation to CO2 & H2O
pyrolysers or gasifiers however these
are both subsets of the overall Biomass Reducing CO2 Oxidation
to CO to CO2
combustion process; hence they can
Char
all be modelled with a comprehensive Ash
combustion model. It is also relevant Char reaction
to point out that the total energy here reduces NOx
available is independent of the
process stages used, however the
conversion of the energy into
electricity may be more efficient for
certain processes. Figure 1
The design and operation of plants involving these processes requires the development of a
mathematical model that is preferably based on the fundamental principles of physics and
chemistry. The reliability of the model must be checked experimentally and the discussion
below presents studies that have been carried out to achieve these aims.

Mathematical Description of Biomass Combustion on a Packed Bed


A packed bed is an assembly of individual particles and consists of a solid phase (the
particles) and a gas phase (gases flowing through the gaps between the particles). Theoretical
calculation of the mass and heat transfer inside a packed bed is made complicated by three
major factors:
1) The temperature profile inside a single particle is highly 3 dimensional (not 1-D in
respect of the radial distance), especially for very thermally thick particles (the Biot
number = hsdp/p >>1);
2) The number of individual particles in a bed is huge prohibiting calculations based on
solving for individual particles; and
3) Lack of models calculating the mixing rate between the under-grate combustion air
and the volatile gases released from solid devolatilization.

2
Other uncertainties include irregularity of the particle shapes, the process rate and channelling
in the bed, etc. For a moving bed, theoretical calculations are further complicated by the
movement and mixing of individual particles which are governed by friction between
particles (dependent on particle shape and orientation), gravity, type of the grate and its
moving pattern during bed operation.
To make the mathematical calculation possible for a packed bed, it is assumed that the major
bed properties, i.e., temperatures of gas and solid phases inside the bed, gas compositions (O2,
H2, CO, CO2, etc.) and solid compositions (moisture, volatiles, fixed carbon and ash) can be
described pseudo one-dimensionally as functions of bed height. It is also assumed that the
bed can be treated as a porous medium where mass and heat transfer take place between the
solid and gas phases and the shape of the particle is spherical (the surface-volume averaged
diameter is used). Under such assumptions, the individual bed processes (moisture
evaporation, devolatilisation and char burning) can be viewed taking place layer by layer,
from the bed top to the bottom. Employing numerical methodology, the whole bed is divided
into many small cells along the bed height and inside each cell the major bed parameters are
assumed uniform. One benefit of this approach is that by reducing the cell size (hence
increasing the cell number), calculation can be made on a size-scale much smaller than the
fuel particles. This means that the non-isothermal behaviour of the single particles can be
accounted for to some extent.
Transport equations for gas and solid phases. Peters (1995) has summarised the basic
governing equations for both the gas and solid phases in a moving bed. For a stationary bed,
the gas-phase equations can be written as follows:

Gas continuity:
( )
g
+
(
g Vg ) = Ssg (1)
t x
where Vg is the gas velocity and x the coordinate along the bed height (x=0 at the bed
bottom). The source term Ssg is the conversion rate from solid to gas due to moisture
evaporation, devolatilisation and char combustion.
Gaseous species transport:
(
g Yig ) + ( g Vg Yig ) = ( )
g Vg Yig
D ig + Syig (2)
t x x x
Yig represents mass fractions of individual species (e.g. H2, H2O, CO, CO2, CmHnOl, ). The
source term Syigaccounts for mass sources of the individual species during evaporation,
devolatilization and the combustion of volatile gases and char.
The fluid dispersion coefficient Dig is considered to consist of diffusion and turbulent
contributions and is given by the following equation (Wakao & Kaguei 1982)
Dig = E0 + 0.5dpVg (3)
where E0 is the effective diffusion coefficient.
Gas-phase energy conservation:
(
g H g ) + ( g Vg H g ) = Tg
g + Sa hs (Ts - Tg ) + Qh (4)
t x x x

where Hg represents gas enthalpy, g the thermal dispersion coefficient, and Qh the heat gain
of the gas phase due to combustion. The thermal dispersion coefficient g consists of
diffusion and turbulent contributions in a similar way as species dispersion, and can be
expressed as [20]:
g = 0 + 0.5dpVg g Cpg (5)

3
Where 0 is the effective thermal diffusion coefficient.
The equations for the solid phase are:

Solid continuity:
(
(1 ) p ) +
(
(1 ) p Vs ) = - Ssg (6)
t x
where p is the particle density and Vs solid velocity due to the downward movement of the
bed caused by mass loss.
Conservation of solid-phase species:
(
(1 ) p Yis ) +
(
(1 ) p Vs Yis ) = - Syis (7)
t x
where Yis represents mass fractions of particle compositions (moisture, volatile, fixed carbon
and ash) and Syis the source term. Syis accounts for the loss of the individual components
(moisture, volatile, fixed carbon and ash) during evaporation, devolatilisation and char
combustion.
The energy equation for the solid-phase is:
(
(1 ) p H s ) + ((1 ) p Vs H s ) = Ts
s + Sa hs (Tg - Ts ) +
q r
+ Qsh (8)
t x
x x x

where Hs presents the solid-phase enthalpy, s is the effective thermal conductivity of the
solid bed, and qr denotes the radiative heat flux. The source term Qsh accounts for the heat
generation due to heterogeneous combustion.

Radiation Heat Transfer in the Bed. Radiation is the major mechanism of heat transfer
between solid particles in a packed bed, and a proper model has to be developed to simulate
the process. The already widely used flux model (Smoot & Pratt 1979) for gaseous and
entrained-flow combustion is the first choice, although development of a more appropriate
model is needed in the future. A two-flux radiation model is presented in the following:

dI +x 1 1
= - (ka + ks) Ix+ + ka Eb + ks ( Ix+ + Ix- ) (9a)
dx 2 2

dI x 1 1
= - (ka + ks) Ix- + ka Eb + ks ( Ix+ + Ix - ) (9b)
dx 2 2
where Ix+, Ix-, represent the two radiation intensities. ka and ks denote the absorption and
scattering coefficients respectively. Eb is black-body radiation.
ks is assumed zero as the first approximation, and ka is taken as ( Shin & Choi, 2000)
1
ka = ln( ) (10)
dp
More details of the model description can be found in Yang et als work (2001, 2002, and
2003).

Mixing Rate of the Under-grate Air with Volatile Combustible Gases Released from Solids
Gaseous fuels released from the devolatilization process have first to mix with the
surrounding air before their combustion can take place. Obviously the burning of the volatile
hydrocarbon gases is limited not only by the reaction kinetics (temperature dependent) but
also by the mixing-rate of the gaseous fuel with the under-fire air. The mixing rate inside the

4
bed is assumed to be proportional to energy loss (pressure drop) through the bed and by
recalling the Ergun equations can be expressed as:

D g (1 )2 3 Vg (1 )1 3 C CO 2
R mix = Cmix g { 150 + 1.75 } min{ fuel , } (11)
d 2p d p Sfuel SO 2

here Cmix is an empirical constant, Dg the molecular diffusivity of the combustion air, Vg the
air velocity, dp the particle diameter, the local void fraction of the bed, C the mass fractions
of the gaseous reactants and S their stoichiometric coefficients in the reaction.
The actual reaction rates of volatile species are taken as the minimum of the temperature-
dependent kinetic rates and their mixing-rates with oxygen:
R = Min[Rkinetic, Rmix] (12)

FLIC CODE PREDICTIONS OF BIOMASS COMBUSTION AND COMPARISON TO


EXPERIMENTS:
Predictions in a large scale moving bed. Based on the mathematical model described above,
a special computer code, FLIC (Fluid Dynamics of Incinerator Combustion) was written to
solve simultaneously the various parallel equations for solid fuel combustion and gasification
in a packed bed. Figure 2 shows some of the results for a large-scale moving bed. Figure 2a)
shows the temperature profile for the whole bed and it can be seen that ignition occurs about 2
meters from the fuel entrance. The reaction front is 150 300 mm (or 2.5 5.0 times the
particle diameters) below the bed top. After ignition, intensive burning occurs above the bed
and long flame tongues are observed. Further on along the bed length, the flame front travels
further into the bed and an increasing proportion of the burning processes occur inside the
bed. The flame front reaches the grate
about two thirds of the bed length and
remaining combustion (mainly char
burnout) continues for a further two
meters until the whole combustion is
completed. Figure 2b) shows the
moisture concentration inside the
burning bed. It is seen that moisture
evaporation occurs in a thin layer. This
is because the high radiation flux from
the flame front quickly heats up the wet
layers inside the particles.

Figure 2a)
Figure 2b) Figure 2c)

All the moisture in the bed is evaporated at about half of the whole bed length.

5
Figure 2c) shows the volatile matter profile inside the bed. The devolatilisation process
occurs also within a narrow layer and all the volatile material in the bed is also released at
about half of the bed length. Figure 2d) shows the individual process rates along the bed
length. Moisture evaporation occurs as soon as the fuel is pushed into the burning chamber
and exposed to freeboard flame radiation. However, it is not until at 2 meters from the fuel
entrance that the top part of the bed has dried out and the temperature is raised above the

1600
450
I II II Test one Test two:T
1400
400 Moisture Tmax
evaporation Volatile rele 1200
Process rates, kg/m2.hr

350
300 1000
250
800
200
Char burn-
600
150
100 400
50
200 Tmin
0
0 2 4 6 8 0
0 2 4 6 8 10
Distance along bed length, m
Distance along bed length, m

Figure 2e). In-bed measurement of temperature


Figure 2d). FLIC prediction of individual using an electronic device and FLIC calculation of
process rates inside a large-scale moving bed. the maximum and minimum local-bed temperatures
Initial fuel moisture 36%. Particle size: 60mm. from the bed top to a distance of 250mm underneath
Initial bed height 1050 mm. in the full-scale moving bed

threshold for volatile matter release. The subsequent volatile release is intense due to the
freshness of the fuel. After that, the devolatilisation rate drops to a stable level. At 6 meters
from the fuel entrance, the volatile release rate rises again, due to the dried-out nature of the
local bed leading to the raised bed temperature, before finally being reduced to zero. The
char starts to burn after the initial intensive release of the volatile matter from the solids and
the burning rates undergo a slow increase as the fuel moves along the bed. At 6 m from the
fuel entrance, the char burning rate rises sharply due to the total release of the volatile matter
and therefore the full access to the under-grate air supply that is no longer consumed by
combustion of volatile gases. The whole combustion process is complete at 7.5 meters along
the bed length. Figure 2d) also indicates that the whole burning process is divided into three
stages: I the ignition stage; II the main stage; and III the final char burnout stage.
The Ball Instrument. To investigate the local combustion behaviour in a full-scale moving
grate incinerator (burning largely biomass) for a range of operating conditions, we have
developed a 'ball instrument' that passes
through the bed with the waste. This small
ball instrument contains instruments and
withstands temperatures up to 1000oC for well The
over an hour (Yang et al. 2001). This novel Ball Instrument
device passes through industrial moving grate passes through
furnaces with the fuel and records parameters the burning
such as oxygen concentration, vibration and bed
several temperatures onto a computer memory
chip. The ball is recovered from the ash pit and
the information is then downloaded onto an
Excel spreadsheet for detailed analysis and
verification of the FLIC predictions.
FLIC code predictions generally compare very well with the data, but the measured
temperature within the bed again shows that the process is dominated by many violent
transient fluctuations from 300C to 1000C along the bed. It was deduced that these
fluctuations were due to the formation and collapse of channels within the bed. A vibration
transducer installed within the ball instrument confirmed this concept since the temperature

6
fluctuations coincided with the mechanical disturbances. The significance of channelling is
also apparent in the jets of flame that can be seen above the bed in incinerator plants. Clearly,
it is not sufficient to model only the mean values of parameters in the bed, and unsteady
aspects of the process have had to be attacked. To investigate this phenomenon, the next
stage has been the successful modelling of the formation of gas flow channels by the random
packing of particles (discussed below), including the effect on the flow distribution caused by
the grate design. Further analysis of the gas combustion within the channels shows that this is
limited by gas phase mixing, and it has proved possible to confirm the predicted height of the
flames above the bed on full-scale plants.
The measured and simulated temperature profiles along the bed length are shown above in
Figure 2e). Data from two runs with same operating conditions are presented. During the
first run only one thermocouple was used
for measurement of the local bed
temperature and during the second run, two 25
Test two: O2
thermocouples were used. These two

O2 concentration, vol% (dry)


20
thermocouples extended from the pair of O2max
sidewalls of instrument opposite each other 15
and were around 200mm apart. The
measurement shows a sharp temperature 10
rise at a distance of 2.0m from the fuel O2min

entrance, indicating the start of fuel ignition 5

that was then followed by a series of violent


0
fluctuations of the measured temperature. 0 2 4 6 8
The position of bed ignition was consistent Distance along bed length, m
with the visual observation from the
viewing ports during the tests in which no Figure 2f). In-bed measurement of O2 concentration
flames were seen for the first 1.5 - 2 m of
the bed length. Figure 2f) shows the measured local O2 concentration profile inside the bed as
the electronic device tumbled along with burning wastes and also the simulated O2
concentration profile. Measurement shows that oxygen began to fall at a position of 1.7 m
from the fuel entrance. It then fluctuated between 0% and 14% for a significant portion of the
bed length (2m 5m) before settling at a more or less stable level (around 4%) after 5 m from
the fuel entrance. Violent fluctuations in both the measured temperature and O2 level are due
to three factors:
1) Constant changing of the probe positions as the electronic device tumbled along the
bed so that the probe tips could be either, out of or inside, the reaction zone which is
about 150 mm to 300 mm deep from the bed top (according to FLIC simulation);
2) Channel formation and destruction in the bed; and
3) Diversity of the fuel properties causing uneven local combustion.
The Effect of under-grate air flow rate combustion or gasification
A one-dimensional pot burner has
been used to provide data for use in W eighing
S cale
Batch Incinerator Expt.
FLIC and to confirm the results of To
S tack

FLIC calculations using various S am pling


S econdary
air

biomass materials. The apparatus is point

15
illustrated in Figure 3. 14
13
S econdary A ir
N ozzle
12
The FLIC code has been used to 11
10 B urner
D ata
calculate the effect of under-grate A cquisition
S ystem
9
8
7
W aste
6
airflow rate in a wide range and the 5
4
R otam eter
3 P rim ary
results are shown in Figure 4. All the 2
1 A ir

calculations were based on fuel G rate A ir


P reheater
C om pressed
air

analysis (listed in Table 1) used in this F igure 3 S chem atic D iagram of the E xperim ental Facility

stationary bench-top reactor (Yang et


al. 2002).

7
Table 1 Fuel Characteristics
Moisture Ash Volatile Fixed carbon C H O LCV Size
No.
wt% wt% wt% wt% wt% wt% wt% MJ/kg mm
1 10 4.3 74 11.7 39.8 4.7 41.2 13.4 12
2 20 3.9 65.7 10.4 35.3 4.1 36.7 11.6 12
3 30 3.5 57.4 9.1 30.7 3.6 32.2 9.85 12
4 40 3.2 49.1 7.7 26.3 3.0 27.5 9.0 12
5 50 2.7 41.0 6.3 21.8 2.5 23.0 8.1 12

Burning rate,dry
Figure 4 shows the burning rate
Moisture content: vs. under-grate airflow rate at
350 fuel moisture = 10% 10%-10mm (WC,Gort different moisture levels, and
1995)
comparison was made between the
300 30%-10mm (WC,Gort
1995) FLIC calculations and
250 40.3% (FW, Thunman experimental data. The airflow
2001)
200
rate spans a range from 0.03 to 0.6
kg/m2.hr

10% this work


20% kg/m2s without preheat (around
150 30% this work 15C) and the moisture level
100
40% this work
covers from 10% to 50% on a wet
30%
40% basis. At each moisture level,
50 50% 50% this work there was a characteristic airflow
0 rate where the burning rate
0 0.2 0.4 0.6 0.8
Under-grate air mass flow, kg/m2.s
reached a maximum and this
characteristic flow rate increases
with decreasing moisture level in
Figure 4. Burning rate vs. airflow rate at different moisture levels. the fuel.
Lines calculations using FLIC; Symbols experimental data from
this work and references. WC wood chips; FW forest waste.

Figure 5 shows the relationship between combustion stoichiometry and under-grate air
flowrate at different moisture levels in the bio-fuel. It is seen that the higher the airflow rate
and wetter the fuel, the richer the
Air to fuel st oichiometric rat io combustion will become. For
M oisture :
2.5 each moisture level, there is a
8.3%(M arie
50% 40% 2000) critical flow rate below which
2.0
30%
20%
30%(Gort
1995 )
the combustion becomes a
1.5
combustion
10% 10% this gasification process (overall air
work
30%this to fuel stoichiometric ratio <1)
1.0 work where a net production of
40%this
0.5 Gasificat ion work combustible gases (CO, CH4,
50%t his
work
etc.) would result. This critical
0.0 flow rate decreases with increase
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Primary air mass f low, kg/m2.s
in the fuel moisture level, i.e.,
wet fuels tend to be combusted
and dry fuels gasified if similar
Figure 5. Combustion stoichiometry vs. airflow rate at different conditions are applied.
moisture levels. Lines calculations using FLIC; Symbols
experimental data from this work and references.

8
Figure 6 shows the peak flame
temperature vs. airflow rate at Peak temperature in the bed

different moisture levels. For 1700


very wet fuels (50% moisture, 1600
for example), the flame 1500
temperature rises quickly with 1400 Moisture content:
increasing airflow and peaks at a 1300
8.3% (Marie et al. 2000)

K
10%
certain critical point. Further 1200 20%
increasing the airflow rate would 1100
30%
decrease the flame temperature 40%
1000 50%
and eventually lead to flame
900
extinction. But for very dry 0 0.1 0.2 0.3 0.4 0.5 0.6
fuels (10% moisture, for Air mass flow, kg/m2.s
example), this critical flow rate
is far higher and so is the Figure 6. Peak temperature vs. air flow rate at different moisture
levels. Line and symbol calculations using FLIC; Symbols
maximum flame temperature Rnnbck Marie et al. 2000).
obtainable. It is also seen that at
lower airflow rate, wet fuels can
produce a higher temperature flame than a dry fuel. This is because the combustion becomes
much more fuel rich (gasification) for dry fuels at low air rates.
Figure 7 shows the calculated CO
CO emission from the bed top concentration in the flue gases
40 exiting from the bed top. As the
35
Moisture content: under-grate air flow decreases, CO
30
10% increases in the flue gases as the
CO (vol%, dry)

20% process becomes air lean and


25
30%
20 shifts from combustion to
40%
15 50%
gasification. Wet fuels produce
10 less CO at the bed top as they
5 favor fuel-lean combustion in the
0 bed.
0 0.1 0.2 0.3 0.4 0.5 0.6
Primary air mass flow, kg/m2.s

Figure 7. CO emission at the bed top vs. airflow rate at different


moisture levels.

PYROLYSIS OF WASTE/BIOMASS AND GENERATION OF STORABLE FUEL


Biomass pyrolysis is one of the thermal energy recovery processes, which has the potential to
generate oil, char and gas products. It is achieved by heating the material in a closed
container in the absence of an oxidizing agent. This decomposes the organic material to
release gas and liquid products with solid char that can be used to generate heat or power.
The process parameters which have the major influence on the products are the pyrolysis
temperature, heating rate, particle size and retort atmosphere (Williams and Besler, 1996;
Beis et al, 2002). The process conditions of pyrolysis can be optimized to maximize the
production of either: pyrolytic oil, char, or gas, all of which have a potential use as fuels. The
carbon char remaining is a valuable fuel that can be easily separated whilst it is still hot from
any metal and stones that were originally present in the bio-waste. It is proposed that the char
and oil can be produced locally in a small-scale simple unit that uses the gas produced to heat
the pyrolyser. The char and oil are than compact fuels that can be transported economically
to a central power station where the advantages of scale can be employed for the efficient and
economic generation of electricity, possibly using gasification as discussed later.
The pyrolysis oils are already being studied elsewhere and the main objective of this study is
to address the relevant aspects of production, characterisation and use of a storable char fuel
derived from the pyrolysis of biomass materials. This will thus contribute to national

9
sustainability targets. A series of pyrolysis tests were conducted initially using wood cubes in
the temperature range of 350C - 700C with a heating rate of 10oC/min. The pyrolytic
products from the tests were then analysed.
Experimental Method The reactor system built was a batch type packed bed pyrolysis unit
shown in Figure 8. It consisted of a reactor in a temperature-controlled furnace followed by
two liquid condenser/traps. The stainless-steel reactor, 12.5cm diameter 50 cm high, is
placed inside a furnace whose inner temperature (To) was controlled by a temperature
controller. Nitrogen gas was supplied from below the reactor to purge the volatile gases
released from the sample during pyrolysis. The volatile gases and nitrogen leaving the reactor
passed through two water-cooled condensers to separate oil vapour from the gas stream. The
oil was collected in a disposable plastic container at the bottom of each trap.
The concentrations of CO, CO2 and O2 in the off-gas after the condensers were monitored by
gas analysers and recorded by the data logger. The flow rate of the sampling gas was l.0
l/min. Gas samples were also taken into glass bottles for further analysis of their chemical
composition using an off-line gas chromatograph.

EXTRACT
SAMPLING BOTTLE
GAS ANALYSER

T4 DATA LOGGER
(CO, CO2, O2)

THERMOCOUPLES
WATER
CONDENSER

FURNACE
T1
REACTOR

T2
BED SECTION
OF SAMPLES
T3
TEMPERATURE
T5 To CONTROLLER

NITROGEN
OIL

OIL CONTAINERS

Figure 8. The pyrolyser set-up

Pinewood cubes of
Sample
size 2cm
Moisture 8.86%,
Volatile matter
Proximate 78.86%,
analysis Fixed carbon
12.08%,
Ash 0.20%
Ultimate C 47.9%, H 6.2%,
analysis O 38.3%
Lower heating
17.8 MJ/kg (dry)
value

Table 2. Properties of the wood sample

10
At the beginning of a pyrolysis experiment, the reactor was charged with feed material,
typically 100 - 400g in weight, and was placed inside the furnace. Then, the furnace was
heated up to the final temperature at the given heating rate with a fixed flow rate of nitrogen.
Once the furnace attained the set value of To, it was maintained for 2 hours to allow sufficient
time to complete pyrolysis.
Characterisation of Wood. The waste sample at the reported initial stage of the experiments
consisted of cubes of pinewood of size 2cm. Table 2 shows the properties of this material
acquired from the proximate, ultimate and calorific value analyses.
100

80
Weight (%)

60

40

20

-dX/dTWOOD
0
0 100 200 300 400 500 600 700

Temperature (oC)
Figure 9. TGA curve for pine sawdust under
nitrogen (heating rate: 10oC/min)

Figure 9 shows the thermo-gravimetric analysis (TGA) for pine sawdust. The pyrolysis
commenced around 250C, and the rate of pyrolysis increased slowly and reached a peak at
386oC. By 400C, the sample had evolved 67% of its original mass by pyrolysis. The rate of
pyrolysis slowed down significantly above 410C.
Experimental Results. Figure 10 shows the solid and liquid products from pyrolysis from a
test with a final temperature of 500oC and the heating rate was 10oC/min. The size of the char
cubes was around 1.5cm, which was 42% of the original wood cube in volume. The mass
yield of the liquid was 45.8% in this case and consisted of a mixture of black greasy oil and
translucent solution. The other pyrolytic product was the off gas composed of CO, CO2, H2,
CH4 and other hydrocarbons. When the temperature reached 500oC, the gas concentrations
were CO 40.4%, CO2 38.9%, CH4 15.8% and H2 4.9%.

Wood cubes Char


size: 2cm 24.5%wt

Liquid
collected
in the traps
45.8%wt

Inside the reactor Gas


after pyrolysis CO, CO2, H2, H2O, CH4,
Hydrocarbons
29.7%wt
Figure 10. The solid and liquid products from pyrolysis

The mass yield, higher heating value (HHV) and corresponding energy yield of char from
various final temperatures are plotted in Figure 11. The energy yield was calculated simply

11
by multiplying the calorific value with the mass yield. The mass yield of char was 33% at
350oC, and decreased with increasing the final temperature. The decrease in the mass yield
over 500oC slowed down. However, the HHV of char increased with increasing temperature,
from 30.9MJ/kg at 350oC to 33.3MJ/kg at 700oC. Since the HHV of char was much higher
than that of the original wood, the energy yield ranged from 49% - 33%.
60 40

Energy Yield Calorific Value 2.0


50 35

Calorific Value (MJ/kg)


wood
40 30
1.5
Yield, %

30 25
Mass Yield

H/C ratio
1.0
20 20 350oC
400oC
10 15 500oC
0.5
600oC
0 10 700oC
300 400 500 600 700 800
o 0.0
Final Temperature ( C) 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Figure 11. Mass yield, energy yields and O/C ratio
calorific value of char from wood (heating Figure 12. Van Krevelen diagram: H/C and
rate=10oC/min) O/C ratios of char (heating rate=10oC/min)

Figure 12 shows a plot of H/C versus O/C ratios, also known as the Van Krevelen diagram.
The char became more carbonaceous at high temperatures. The carbon content in the char at
350oC was 77%, which was 53% of the original carbon in the wood sample. It became 89%
at 700oC, which was 40% of the original carbon in the wood sample. At the final temperature
of 500oC, the carbon content of char was 83.1% (43% of original carbon), while the hydrogen
content was 3.8% (15% of original hydrogen). The decrease in H/C and O/C ratios was close
to a linear relationship.
100 36
Higher Heating Value (MJ/kg)

80
HHV 32
VM
Weight (%)

60 28
FC

40 24

20 20

0 16
350 400 500 600 700 B A
o
Final Temperature ( C)

Figure 13. Proximate analysis of char (heating rate=10oC/min)


(VM: Volatile Matter, FC: Fixed Carbon, B: Bituminous coal, A: Anthracite, Source of the
coal results: Phyllis Database)

Figure 13 shows the volatile matter (VM) and fixed carbon (FC) contents measured from the
proximate analysis of char along with its calorific value. The proportion of FC increased with
temperature rise, which is the inherent nature of pyrolysis. The chars from 400C and 600C
showed similar FC/VM ratios to bituminous coal and anthracite, respectively. Thus the
calorific values of chars were as high as that of bituminous coal and anthracite.
The main conclusions drawn from this phase of the work are:
i) The mass yield of char from wood cubes decreased from 33.0% to 21.5% with
increasing final temperature from 350oC to 700oC at the heating rate of 10oC/min.
ii) The char became more carbonaceous at higher temperatures, and the decrease of
H and O in char showed a linear relationship between H/C and O/C ratios.

12
iii) The calorific value of char was over 30 MJ/kg, which was as high as coal.
As discussed above, in addition to the well-reported value of the pyrolysis liquids as a fuel
oil, the use of the storable char as feedstock for a gasifier required investigation. Its
behaviour was therefore studied in our unique gasification system.

ULTRA SUPER-HEATED STEAM GASIFICATION


Introduction. The conversion efficiency of solid fuel via steam raised in a boiler into
electricity using steam turbines is only about 20% to 45% depending on the temperature and
pressure of the steam. Gasification of solid fuels for IGCC systems produces gas to run gas
turbines that are used to produce power together with steam turbines. This enables enhanced
efficiency in the conversion of energy from solid fuels and the overall efficiency of a
conventional integrated gasification combined cycle is about 51%. In this case, the heat
required for the gasification reaction is generated by the reaction of about 15% of the fuel
with oxygen. A new process is proposed that uses low-grade steam from an incinerator as the
starting point for an ultra superheated steam gasification system for solid fuels. The overall
efficiency of the proposed process is about 60%, which is superior to the sum of the separate
processes of electricity from a steam boiler together with an independent biomass
gasification-based electricity generator.
When used as an embedded CHP
Steam
Ultra Superheated Steam
for Sustainable Processes
system, the energy complex can
produce power and heat from biomass
(and coal) for a sustainable city at an
Steam at 1600C efficiency of about 85%. If the plant is
installed in a city, the biomass can be
derived from waste, thus disposing of
waste that cannot be effectively reused
or recycled.
The aim of this aspect of the research
Thermocouple project is the development of the novel
gasification process that utilises ultra-
superheated steam (Figure 14).

Figure 14
Experimental investigation. The new gasification concept offers the prospect of achieving
the goal of high efficiency power generation from various biomass (or coal) energy sources
by utilizing Ultra Superheated Steam (USS). The concept first reduces the biomass including
wastes such as wood waste (and coal) to a fine powder that can be reacted in a small
residence time and therefore a small volume. The process uses low-grade steam from sources
such as waste incineration, process cooling, and local industry, and then enhances it to a
temperature above 1600C. This is achieved by adding oxygen to the steam to form "artificial
air"; gas is then burned in this artificial air to produce an ultra-superheated steam flame. The
biomass (or coal) is injected coaxially into the high temperature steam flame where it reacts in
about one second to produce a gas that is free of tar and consists largely of H2, CO and CH4.
Significantly, the high temperature steam provides the enthalpy for the endothermic
gasification reactions. This feature ensures that any ash is cooled below its melting point
before it arrives at the reactor walls. This is important since the high alkali content of
biomass ash results in a low melting point and hence a high tendency to form slag.
The USS gasifier can be envisaged as being composed of two distinct main parts, namely:
1. The USS generator (a conventional burner where the gasification also commences).
2. The gasifier (which is the shell where the reactions proceed).

The USS Burner. The North American Manufacturing Company manufacture the burner
selected for the generation of USS and subsequent USS gasification (Figure 15). Its design

13
capacities are a maximum airflow of 0.0322 m3/s and a heat output of 120 kW. Its
construction features include air and gas inlets into the burner and a standard quarl that is
about 230 mm long.
This burner was chosen because, being manufactured for dual fuel operation it could be easily
adapted for the USS gasification by replacing the liquid fuel pipe with one suitable for the
supply of granular material for the gasification of solids, or by just using the burner as
supplied for the gasification of liquids such those from pyrolysis, or slurry material made
from these oils combined with char.
This burner was adapted to feed in particulate materials by removing the liquid fuel inlet pipe,
and replacing it with a pipe with a wider bore, connected to the funnel and vibrating feeder
system as shown in Figure 15.
The USS Gasifier. The rational design of a gasifier requires the best possible understanding
of the fundamental chemical and physical
processes that occur: Pressure is one of the
operating variables of interest in the
investigation of USS gasification. In the
gasifier design stage, high pressure USS
gasification was considered but the extent of
the initial pressure was limited by the rapid
construction and operation feasibilities. This
then limited the pressure operation of the
gasifier to near atmospheric for the
preliminary tests, where the material to be
gasified is fed into the gasifier by gravity.
The gasification chamber is a cylindrical mild
steel shell lined with a 50mm thick fused
alumina based castable refractory with high
abrasion resistance. The refractory inside
diameter = 285 mm. A catch pot at the bottom
collects ash or slag. The USS gasification was
carried out using methane as the gaseous fuel
for the generation of USS and propane was
used to fire the pilot burner. The yield of gas
was measured by collecting samples for GC
analysis and also evaluating the ash content.
A photograph of the equipment is shown in
Figure 15. Figure 15
The theoretical calculations were carried out to estimate the gas yield from USS gasification.
The Boudouard, water-gas and hydrogasification reactions occur simultaneously. The
predicted yield at equilibrium was used for carbon gasification using steam at 1 atm and
2131K (which is the adiabatic flame temperature for a stoichiometric USS flame). The results
are:
Volume fraction of CO = 54.8%
Volume fraction of CO2 = 0.0%
Volume fraction of H2 = 45.2%
Volume fraction of H2O = 0.001%
Volume fraction of CH4 = 0.007%
Chemical equilibrium calculations based on minimising Gibbs free energy are shown in the
figure 16. These show higher levels of CO2 as observed in the tests.

14
Figure 16

THE INTEGRATION OF FLIC WITH FLUENT


In order to design industrial plant based on the various processes discussed above, integrated
process modelling techniques are required. These are include advanced procedures such as
the recognition and modelling of the channels that form in packed beds of particles such as in
moving grate combustion systems. Such an
assortment of particles is illustrated in Figure 17.
The flow distribution through such a bed takes
place along channels and the results of
calculations (that have been verified by
measurements) is illustrated in Figures 18 a) and
b). These show the calculated flow in a vertical
section and that measured across the bed surface
respectively. The formation of such burning
channels can also be seen clearly by looking at
the flames on the surface of a burning bed of
coal or other particles. This phenomenon is also
Figure 17 included in the FLIC model.

Measurement of channelling in a
Simulation of channel formation
pot bed without combustion
in a non-
non-burning bed
V/V0

Grate causes initial uniform flow

Figure 18 a) Figure 18 b)
FLIC thus provides a boundary condition for the FLUENT CFD calculation of the flow in the
freeboard above the burning bed. However, FLIC requires information on the radiation heat
transfer from the freeboard. Thus the calculation procedure must be performed by the

15
simultaneous (iterative) solution of both FLIC and FLUENT. This procedure for the biomass
combustion in a large incinerator is illustrated in Figures 19 a). Figure 19 b) shows the
temperature distribution through the plant. Figure 19c) shows that a flat heat transfer
distribution used to initiate the iterative procedure leads to convergence after about four
cycles. Figure 20 shows the converged results of the combined reacting flow calculation.

F L U E N T /F L IC m o d e l in te g ra t io n

W ATER W ALL

W ASTE B O IL E R
FLUENT TUBES
( G a s F lo w M o d e l)

M GAS, T GAS, V GAS SECONDARY


A IR
Q RAD

GRATE
F L IC
(M o d e l fo r W a s te B e d )
P R IM A R Y A IR ASH

Figure 19a)

S im u la t io n R e s u lt s
G A S R E LE A SE D A TA G A S F L O W F IE L D
F R O M F L IC FRO M FLUENT

T e m p e ra tu re
Gas Properties

V e lo c it y

D is ta n c e a l o n g b e d
IN C ID E N T R A D IA T IO N
FRO M FLUENT
W A S T E C O M B U S T IO N
Radiation

F R O M F L IC

D is ta n c e a l o n g b e d

Figure 19b)

C a lcu la tio n R esu lts


F L U E N T to F L IC : In c id e n t ra d ia tio n o n th e w a s te b e d
E m itte d g a s a ffe c ts th e ra d ia tio n p ro file sh a p e
(a ) B y g a se o u s e m iss io n , (b ) V ia th e fu rn a ce w a ll,
(c ) B y p a rtic u la te e m issio n .
T h e ra d ia tio n p ro file b e ca m e s ta b le a fte r th e 4 th
ite ra tiv e c a lcu la tio n s o f F L IC a n d F L U E N T
Effective Radiation Temperature (K)

1 6 00 1 6 00

1 4 00 0 .5
Gas Temperature (K)

1 4 00 1 .5
1 2 00 2 .5
1 2 00 3 .5 th
1 0 00
1 0 00
0 (a s s u m e d )
1 800
800
2
3 600
600
4 th u pd a te
400
400
0 2 4 6 8 10 0 2 4 6 8 10
D is ta nc e a lo n g b e d le n g th(m ) D is ta n c e a lo n g b e d le n g th (m )

Figure 19c)

16
FLUENT/FLIC Results: Gas Concentrations
Note the active gaseous reactions by fresh
oxygen from the secondary air

O2 MASS FRACTION CO2 MASS FRACTION

CO MASS FRACTION O2 MASS FRACTION TEMPERATURE

Figure 20

CONCLUSION
The combustion/pyrolysis/gasification of biomass on both static and moving beds has been
investigated. The fundamental processes have been studied and an understanding has been
gained of the complexities involved. New mathematical models of these phenomena now
give a basis for integrated modeling that can provide design data for industrial exploitation of
energy from biomass.
Acknowledgments:
The authors would like to thank the following organizations for their financial and technical
support for the above research programmes: UK Engineering and Physical Science Research
Council (EPSRC), Onyx Environmental trust and the UK Incineration Industry.
Nomenclature
Av pre-exponent factor in devolatilization rate, s-1
C constant; molar fractions of species (fuel, oxygen)
Cfuel fuel concentration, kg/m3
Cpg specific heat capacity of the gas mixture, J/(kg K)
Cmix mixing-rate constant
Dg molecular diffusion coefficient of volatile hydrocarbons in air, m2/s
Dig dispersion coefficients of the species Yi, m2/s
dp particle diameter, m
Dr in-flow dispersion coefficient in bed, m2/s
Ds particle mixing coefficient due to random movements of particles in the bed, m2/s
Ev activation energy in devolatilization rate, J/kmol
Hg gas enthalpy, J/kg
Hs solid-phase enthalpy, J/kg
hs convective heat transfer coefficient between solid and gas, W/m2K
kv rate constant of devolatilization, s-1

17
Qcr heat absorbed by the solids, W/m3
Qh heat loss/gain of the gases, W/m3
Qsh thermal source term for solid phase, W/m3
qr radiative heat flux, W/m2
R mix mixing-rate of gaseous phase in the bed, kg/m3s
S stoichiometric coefficients in reactions
Sa particle surface area, m2
Ssg conversion rate from solid to gases due to evaporation, devolatilisation and char burning,
kg/m3s
Syig mass sources due to evaporation, devolatilization and combustion, kg/m3s
Syis source term, kg/m3s
t time instant, s
Tg gas temperature, K
Ts solid temperature, K
Vg superficial gas velocity (vector), m/s
Yis mass fractions of particle compositions (moisture, volatile, fixed carbon and ash)
s system emissivity
b Boltzmann radiation constant, 5.86 10-8 W/m2K4
void fraction in the bed
g gas density, kg/m3
sb solid bulk density in the bed, kg/m3
g thermal dispersion coefficient, W/mK
s effective thermal conductivity of the solid bed, W/mK
References:
Beis S.H.; Onay O and Kockar O.M., Fixed-bed pyrolysis of safflower seed: influence of pyrolysis
parameters on product yields and compositions, Renewable Energy, 2002, vol. 26, no. 1, pp. 21-32
Gort, R, On the Propagation of a Reaction Front in a Packed Bed: Thermal Conversion of Municipal
Waste and Biomass, Academic Dissertation, University of Twente, 1995.
Peters B, A detailed Model for Devolatilization and Combustion of Waste Material in Packed Beds, 3rd
European Conference on Industrial Furnaces and Boilers (INFUB), Lisbon, Portugal, 18 21 April,
1995.
Pyle, D. L. and Zaror, C. A., Heat transfer and kinetics in the low temperature pyrolysis of solids,
Chem Eng Sci, Vol.39, No.1, pp.147-158, 1984.
Shin, D and Choir, S, The combustion of simulated Waste Particles in a Bed, Combustion and Flame,
Vol.121, pp.167-180, 2000.
Smoot, L D and Pratt, D T, Pulverized-coal Combustion and Gasification, Plenum Press, 1979.
Wakao, N and Kaguei, S, Heat and Mass Transfer in Packed Beds, Gorden & Breach Science
Publishers, 1982.
Williams P.T. and Besler S., The influence of temperature and heating rate on the slow pyrolysis of
biomass, Renewable Energy, Vol.7(3), pp.233-250, 1996
Yang Y B, Goodfellow J, Ward D, Gan S, Swithenbank J and Nasserzadeh V, Cutting Wastes From
Municipal Solid Waste Incinerator Plants, Trans IChemE, Vol 81, Part B, pp.143-155, May 2003.
Yang Y B, Nasserzadeh V, Goodfellow J, and Swithenbank J, Simulation of Channel Growth in a
Burning Bed of Solids, Trans IChemE, Vol 81, Part A, pp.221-232, 2003.
Yang Y B, Nasserzadeh V, Goodfellow J, Goh Y R, and Swithenbank J, Parameter Study on the
Incineration of Municipal Solid Waste in Packed Beds, Journal of Institute of Energy, September,
2002.
Yang Y B, Goh Y.R, Zakaria R, Nasserzadeh V, Swithenbank J, Mathematical modelling of MSW
incineration on a travelling bed, Waste Management 22 (2002) 369-380.
Yang Y B, Goodfellow J, Goh Y, Nasserzadeh V and Swithenbank J, Investigation of Channel
Formation Due To Random Packing In a Burning Waste Bed, Trans IChemE, Vol 79, Part B, 2001.

18

You might also like