Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Article

pubs.acs.org/EF

High-Temperature Conversion of SO2 to SO3: Homogeneous


Experiments and Catalytic Effect of Fly Ash from Air and Oxy-fuel
Firing
Lawrence P. Belo,† Liza K. Elliott,† Rohan J. Stanger,† Reinhold Spörl,‡ Kalpit V. Shah,† Jörg Maier,‡
and Terry F. Wall*,†

Chemical Engineering, University of Newcastle, Callaghan, New South Wales 2308, Australia

Institute of Combustion and Power Plant Technology (IFK), University of Stuttgart, 70569 Stuttgart, Germany

ABSTRACT: The reaction of SO2 with fly ash in the presence of O2 and H2O involves a series of reactions that lead to the
formation of SO3 and eventually H2SO4. Homogeneous experiments were conducted to evaluate the effects of the procedural
variables, i.e., temperature, gas concentrations, and residence time, on the post-combustion conversion of SO2 to SO3. The results
were compared to existing global kinetics and found to be dependent upon SO2, O2, residence time, and temperature and
independent of H2O content. For a residence time of 1 s, temperatures of about 900 °C are needed to have an observable
conversion of SO2 to SO3. Literature suggested that the conversion of SO2 to SO3 is dependent upon the iron oxide content of
the fly ash. Experiments using three different fly ash samples from Australian sub-bituminous coals were used to investigate the
catalytic effects of fly ash on SO2 conversion to SO3 at a temperature range of 400−1000 °C. It was observed that fly ash acts as a
catalyst in the formation of SO3, with the largest conversion occurring at 700 °C. A homogeneous reaction at 700 °C, without fly
ash present, converted 0.10% of the available SO2 to SO3. When fly ash was present, the conversion increased to 1.78%. The
catalytic effect accounts for roughly 95% of the total conversion. Average SO3/SO2 conversion values between fly ash derived
from air and oxy-fuel firing and under different flue gas environments were found to be similar.

1. INTRODUCTION in power plants because of the innately low sulfur content of


Australian coals.
For centuries, fossil fuels (coal, oil, and natural gas) have been
The current paper focuses on understanding the mechanisms
the primary source of energy for power generation, and this is
of the conversion of SO2 to SO3 under post-flame conditions in
expected to continue in the distant future. Coal, the cheapest air and oxy-fuel flue gas environments. Belo et al.8 discussed in
fossil fuel, consists mainly of carbon and results in emissions of a previous paper the reaction routes of sulfur in a combustion
carbon dioxide when combusted. A total of 41% of the system. SO2 is formed from the decomposition and oxidation of
approximately 30 gigatonnes of CO2 emissions in 2010 was organic and inorganic sulfur associated in the coal matrix. SO2
contributed by energy generation alone.1 Apart from being the is then converted to SO3 via either homogeneous gas-phase
largest source of greenhouse gases (chiefly CO2), coal-fired reactions or heterogeneous catalytic reactions.8,11 Conversion
power generation also contributes significant amounts of other of SO2 to SO3 may be influenced by the following factors,
pollutants to the environment. Among the major pollutants are namely, SO2 partial pressure,12−15 O2 partial pressure,13−15
SO2 and SO3, collectively known as SOx, which apart from presence or absence of moisture,12−16 presence of catalytically
being a health concern2 plays an important role in the active components in the fly ash (e.g., Fe2O3),12,13,15,16 and
formation of photochemical smog and acid rain.3,4 CO2 capture temperature-residence time profile of the plant.13
and storage (CCS) technologies have been developed to The principal mechanisms for SO3 production in combustion
address this growing concern of greenhouse gas and pollutant systems are via
emissions. One promising CCS solution is oxy-fuel combustion, SO2 + O2 ⇄ SO3 + O (1)
where coal is burned in a mixture of oxygen and recycled flue
gas (RFG) instead of air. The resulting flue gas, rich in CO2, is SO2 + O ( +M) ⇄ SO3 ( +M) (2)
cleaned and sequestered. with eq 2 occurring at flame temperatures. Burdett et al.17
Studies have shown that using recycled flue gas in oxy-fuel stated that, unless there is a high concentration of O atoms
results in not only higher CO2 but also SO2 in the combustion present, the concentration is not high enough to facilitate SO3
zone, which could change the sulfur partitioning in the flue gas, production via eq 2.
resulting in higher SO3 concentrations.5,6 This is of particular Fleig et al.,11 Jorgensen et al.,13 and Cullis and Mulcahy18
importance when no SOx cleaning of the recycle stream is in stressed that, even though photochemical oxidation of SO2 had
place, and SO2 concentrations can be as high as 4 times that of
air combustion.1,7−10 The increased SO2 concentration Received: September 11, 2014
resulting from recycling of the flue gas is challenging in the Revised: October 23, 2014
Australian scenario, where no flue gas desulfurization is applied Published: October 23, 2014

© 2014 American Chemical Society 7243 dx.doi.org/10.1021/ef5020346 | Energy Fuels 2014, 28, 7243−7251
Energy & Fuels Article

Figure 1. Experimental setup for homogeneous and heterogeneous SO2 to SO3 conversion experiments.

been studied in the past, there is limited convincing evidence of atm. A water saturator maintained at a controlled temperature was
an uncatalyzed homogeneous reaction at temperatures below used to introduce water vapor into a N2 or CO2 carrier gas. The
900 °C. However, according to Cullis and Mulcahy18 and Rees amount of water present was measured by a Testo 350XL relative
et al.,19 oxidation of SO2 to SO3 is rather difficult but can be humidity probe. Reactant gases were heated with heating tapes to
approximately 80 °C before entering the reactor. The experiment
achieved catalytically. Cullis and Mulcahy,18 Marier and consisted of two parts: homogeneous conversion reaction with gases
Dibbs,16 Spörl et al.,20 and Jorgensen et al.13 have shown that only and heterogeneous conversion, with the reaction of gases in the
fly ash could act as a catalytic oxidizer to SO2 given the presence of fly ash.
presence of metal oxides in its matrix even at temperatures as During the homogeneous experiments, the quartz tube was empty,
low as 400 °C. One of the best known catalysts is Fe2O3. The except for the feed gas flowing through it, but during the
reaction is as follows:2,16,19 heterogeneous experiments, the quartz tube held a packed bed of fly
ash (∼0.5 g) supported by quartz wool placed in the center of the
1 Fe2O3
isothermal region. The calculated residence time in the isothermal
SO2 + O2 ⎯⎯⎯⎯⎯⎯→ SO3
2 catalyst (3) zone and the calculated contact time with a ∼1 cm thick bed of fly ash
are presented in Table 1.
During oxy-fuel combustion, the increased amount of O2 and
SO2 has the potential of affecting the degree of oxidation of
Table 1. Calculated Residence Time in the Reactor and
SO2 to SO3, which has been reported in several studies.1,6,10
Contact Time with Fly Ash with an Input Gas Flow of 0.5 L/
Limited literature exists on the conversion of SO2 in the
min at 298 K and 1 atm
absence of combustibles. Spörl et al.20 stressed that the ratios
between heterogeneous and homogeneous SO3 formation are temperature residence time at temperature contact time with fly ash
still vague. The presence of SO3 is significant in power plants (°C) (s) (ms)
because increasing amounts of SO3 in the flue gas increase the 400 1.50 60.1
acid dew point (ADP). At temperatures below ∼400 °C,11 SO3 500 1.30 52.3
is highly reactive with water vapor to form H2SO4, reaching 700 1.04 41.6
complete transformation at ∼200 °C via the reaction8,20 900 0.86 34.5
1000 0.79 31.8
SO3(g) + H 2O(g) ⇄ H 2SO4 (g) (4)
Upon cooling to temperatures below the ADP, H 2 SO4 To determine the conversion of SO2 to SO3, a modified condenser
condenses, which is highly corrosive to boiler components based on the principles of the controlled condensation method25−28
(e.g., air heaters).1,13,21−23 To avoid corrosion, it is imperative (CCM) was employed. CCM assumes that, if an excess partial
that all of the components of the boiler be kept and operated at pressure of water vapor is maintained in the reaction zone,16 all SO3 is
temperatures greater than the ADP, which will, in turn, increase converted to H2SO4, which can then be separated from the gas stream
the overall plant efficiency losses associated with heat recovery.1 by condensation. Passing the gas stream through a condenser at
The purpose of this work is to investigate the SO2 to SO3 temperatures below the sulfuric acid dew point (90 °C)16,25,27−29 and
conversion as a function of SO2, O2, and H2O in post-flame above the water dew point30 (generally below 60 °C25,29) allowed for
conditions in the absence of combustibles. The extent of collection of H2SO4.
Quartz wool was placed inside the quartz condenser to provide
catalytic conversion because of fly ash was also investigated. more surface area for condensation. Conversion of SO2 to SO3 in the
gas mixture was determined by the amount of H2SO4 produced during
2. EXPERIMENTAL SECTION each experiment, completed after 60 min. A known quantity of
The experimental rig used in this study on the post-flame conversion distilled deionized (DDI) water was used to flush out the condenser to
of SO2 to SO3 is shown in Figure 1. The setup consisted of a 12 mm collect H2SO4. The H2SO4 concentration was then determined using a
inner diameter (S/V ratio = 3.33 cm−1) quartz tube flow reactor placed Dionex Dx-100 ion chromatogram (IC). To determine the amount of
in an electrically heated horizontal furnace with an isothermal length of quartz wool needed to fully capture the acid condensate, different
25 cm. Quartz was chosen as the reactor material and packing material amounts (0.25, 0.50, 0.75, 1.0, and 1.5 g) of quartz wool were placed
because previous researchers13,24 verified the inert properties of quartz in the condenser to collect prepared concentrations of H2SO4
to SO2 oxidation. Reactant gases (SO2, O2, and N2/CO2) were vaporized within the furnace. The concentration of H2SO4 was
supplied by gas cylinders and controlled by Brooks mass flow approximately a magnitude higher than the equivalent amount of SO3
controllers with a combined flow rate of 0.5 L/min at 298 K and 1 generated in power plants. Figure 2 shows the graph relating the mass

7244 dx.doi.org/10.1021/ef5020346 | Energy Fuels 2014, 28, 7243−7251


Energy & Fuels Article

Figure 2. Quartz wool capture test with 1.0 and 0.1 M H2SO4.
Figure 3. ADP estimations relating SO3 and H2O vapor concen-
of the quartz wool and the recovery (%) of H2SO4 used. From the test, trations simulating conditions in the flue gas using ZareNezhad’s31
it was observed that an average 92.16−99.98% capture was achieved ADP correlation.
with greater than 1.0 g of quartz wool for both high and low H2SO4
concentrations. Hence, this amount of quartz wool was used
throughout the experiments. Homogeneous conversions investigated the effects of gas concen-
The sulfuric acid dew point (ADP) temperature is one of the trations of O2, SO2 and H2O. Heterogeneous conversions were carried
conventional ways of estimating the amount of SO3 and H2O present out in the presence of fly ash heated to the reaction temperature of,
in the flue gas in power plants.28 ZareNezhad’s correlation31 was used i.e., 400, 500, 700, 900, and 1000 °C.
for estimating the ADP, as shown in eq 5 Fly ash samples used in the heterogeneous conversions were fly ash
obtained from a joint Australian−German study evaluating coal
Tdew = 150 + 11.664 ln(pSO ) + 8.1328 ln(pH O ) − 0.383226 behavior in oxy-fuel. Fly ash samples A, B, and C derived from three
3 2
Australian sub-bituminous coals (A, B, and C) were obtained from the
ln(pSO )ln(pH O ) (5) 20 kWth combustion rig located at the Institute of Combustion and
3 2
Power Plant Technology (IFK), University of Stuttgart, Stuttgart,
where Tdew is the sulfuric acid dew point temperature (°C), pH2O is the Germany.9 The rig was operated at 1350 °C, capable of combustion
partial pressure of H2O in the flue gas (mmHg), and pSO3 is the partial investigations with 0.5−3 kg of pulverized fuel/h and having a constant
pressure of SO3 in the flue gas (mmHg). product rate of 11.5 m3 [standard temperature and pressure (STP)]/h
Selection of the inlet gas concentrations (SO2 and H2O) were to maintain comparable residence times for the different modes of
completed using values measured in pulverized fuel (PF) combustion firing, i.e., air, oxy-fuel with partial cleaning, and oxy-fuel full recycling
and predicted for oxy-fuel combustion6,9,32 as a guide (Table 2). without cleaning. The fly ash was obtained from the bag filter of the rig
maintained at an inlet temperature of 225 ± 30 °C and outlet
Table 2. Parameters Used in the Estimation of the Sulfuric temperature of 195 ± 15 °C. Details of the conditions and the rig are
discussed in previous literature.8−10 The X-ray fluorescence (XRF)
Acid Dew Point Temperature data of the fly ash used in the experiments is presented in Table 4. It
mode [SO2] (ppm) [SO3] (ppm) [H2O] (vol %) can be noted that the Fe2O3 data in the table are bold to emphasize
differing values, with sample B being the lowest and sample C being
air 500 0−100 3
the highest.
oxy partial cleaned recycle 1000 0−100 10
oxy uncleaned recycle 2000 0−100 30 3. RESULTS AND DISCUSSION
Experiments were performed to quantify the conversion of SO2
Experiments labeled as “air” represent PF combustion, in which
cleaned air is used as the oxidant. Experimental concentrations of 1000 to SO3 under post-flame conditions using simulated air and
ppm of SO2, 5 vol % O2, and 3 vol % H2O reflect air combustion flue oxy-fuel flue gas at relatively low temperatures of 400−1000 °C.
gas levels using medium−high sulfur coals. Practical oxy-fuel 3.1. Homogeneous SO2 to SO3 Conversion. 3.1.1. Effect
combustion with partial flue gas cleaning (with simulated removal of of the Residence Time. Experiments with varying gas velocities
nominally 20 vol % H2O, 20 vol % SO2, and 50% Hgtot from flue gas were performed to obtain insight into the effect of the residence
based on theoretical maximum conversion9 before being recycled to time on the system. Gas velocities of 0.5−1.5 L/min (dry, STP)
the furnace) is simulated in the experiments termed here as “oxy were used to obtain residence times of 0.3−0.9 s. Figure 4
partial cleaned recycle”. Oxy-fuel combustion with flue gas recycling presents the effect of the residence time on SO2 conversion to
without cleaning is termed in this study as “oxy uncleaned recycle”. SO3 compared to a previous study completed by Flint and
Assuming 2% conversion of SO2 to SO3 and the input gas
concentrations shown in Table 2, the ADP estimations show that, by Lindsay,24 who used a minimum residence time of 1 s to
changing from air combustion to oxy-fuel combustion, the ADP generate about 0.4% SO3/SO2 conversion. The experimental
increases from 120 to 140 °C, presented in Figure 3, while if the flue results are also compared to a kinetic model produced by
gas was fully recycled without any cleaning, the ADP could potentially Burdett et al.,17 as given by eq 6
increase by 40 °C, from 120 to 160 °C. ZareNezhad’s correlation was
d[SO3] k A[SO2 ][O2 ] (−B / T )
also used to estimate the operating temperature of the controlled = 1 [SO2 ][O2 ] = e
condensation experiment. It estimated that, with even a low SO3 dt RT RT (6)
concentration of 0.5 ppm and water concentrations of about 3 vol %, −1 3 −1
the ADP sits at a temperature of 93 °C, still higher than the condenser where k1 = A exp(−B/T), A = 2.6 (±1.3) × 10 mol cm s , 12

temperature range of 60−90 °C (75 °C was chosen for this study). B = 23 000 ± 1200 K (leading to B/R = 190 ± 10 kJ mol−1),
To investigate the individual kinetic effects of SO2, O2, and H2O in and [SO2], [O2], and [SO3] are partial pressures.17
the post-flame conversion of SO2 to SO3 at temperatures of 400−1000 Increasing SO3/SO2 conversions were observed as the
°C, the following experimental design was used in this study (Table 3). residence time in the reactor was increased. It could be noted
7245 dx.doi.org/10.1021/ef5020346 | Energy Fuels 2014, 28, 7243−7251
Energy & Fuels Article

Table 3. Effect of Varying Gas Concentrations on the Uncatalyzed, Post-flame Conversion of SO2 to SO3 at 900 °C
experiment control variable
(1) SO2 effect [O2] = 5 vol % [SO2] (ppm) = 500, 1000, 1500, and 2000
[H2O] = 3 vol %
(2) O2 effect [SO2] = 1000 ppm [O2] (vol %) = 3, 5, and 10
[H2O] = 3 vol %
(3) H2O effect [SO2] = 1000 ppm [H2O] (vol %) = 3, 4, 7, and 9
[O2] = 5 vol %

Table 4. XRF Data of the Fly Ash Samples Used in the Study
fly ash sample
A B C
(%, dry basis) air oxy air oxy air oxy
SiO2 55.2 52.9 68.7 69 53.8 53
Al2O3 33.3 34 25.2 24.2 25.9 26.5
Fe2O3a 6.55 7.13 1.62 2.16 8.55 8.34
CaO 0.95 1.41 0.73 0.87 4.55 4.66
MgO 0.741 0.98 0.716 0.737 1.558 1.667
Na2O 0.144 0.158 0.25 0.275 0.971 0.979
K2O 0.512 0.518 0.627 0.604 1.11 1.058
TiO2 2.06 2.16 1.77 1.65 1.15 1.13
MnO2 0.097 0.113 0.002 0.008 0.04 0.041
P2O5 0.178 0.275 0.134 0.176 1.364 1.508
SO3 0.132 0.243 0.186 0.28 0.932 0.969
BaO 0.042 0.056 0.067 0.063 0.067 0.072
SrO 0.049 0.069 0.034 0.036 0.094 0.1
total 100 100 100 100 100 100
percent carbon (unburned) <0.1 <0.1 0.106 <0.1 0.134 0.199
a
Fe2O3 XRF values are in bold for emphasis.

concentration on the SO3 output and SO2 to SO3 conversion


are presented in panels a and b of Figure 5. As anticipated, SO3
concentrations were higher when the input SO2 partial pressure
is increased (Figure 5a). However, the data produced a linear
trend with a gradient that is much smaller than that provided by
the Burdett model. The experimental results fall within ±35%
of the predicted values. On the basis of the trend of the data, a
higher disparity would be expected with increasing SO2 input,
i.e., 2000 ppm onward.
Average conversions ranged from 0.48% for 500 ppm of SO2
input to 0.33% for 1000 ppm of SO2, 0.24% for 1500 ppm of
SO2, and 0.25% for 2000 ppm of SO2 (Figure 5b). It can be
Figure 4. Effect of the residence time on the homogeneous conversion seen that, as the input SO2 is increased, a decreasing conversion
of SO2 to SO3. Inlet gas mixture: 1000 ppm of SO2, 5 vol % O2, 3 vol is observed. This behavior was also reported by other
% H2O, and balance N2, at a temperature of 900 °C. investigators;33,34 Schwaemmle et al.33 stated that conversions
at lower concentrations start high but decrease as the input SO2
concentration becomes higher, giving a pseudo-first-order
that, at about 0.3 s of residence time, there was almost reaction. Supporting the claim, Svachula et al.34 stated that an
negligible conversion observed (average 0.03%), whereas at apparent kinetic order in SO2 is higher than 1 for low
∼0.9 s, conversions from 0.14 to 0.36% were observed (average concentrations (0−200 ppm) and decreases to a fractional
0.24%). Longer residence times were not observed because order at higher SO2 inlet concentrations. Conversely, other
they are difficult to achieve as a result of limitations in studies34,35 indicated that a first-order dependence may provide
equipment. Conversions were calculated using the formula a reasonable approximation between 0 and 1000 ppm of SO2.
given by the equation Experiments with varying the O2 concentration were
mol SO3 performed to investigate the effect of O2 on the homogeneous
percent conversion (%) = × 100 uncatalyzed oxidation of SO2. Bayless et al.36 stated that
mol SO2 (7)
previous studies established that at least 1% excess O2 is needed
3.1.2. Effect of Impurity Concentrations in the Flue Gas. for the conversion of SO2 to SO3. Typically, coal-fired plants
Experiments with different concentrations of reactants SO2, O2, use between 3 and 5% excess O2.36 In this study, 3, 5, and 10%
and H2O vapor were also investigated and compared to were used as the O2 levels. The results are plotted in panels a
Burdett’s kinetic model. The effect of the input SO 2 and b of Figure 6. Similar to the effects of SO2, an increasing O2
7246 dx.doi.org/10.1021/ef5020346 | Energy Fuels 2014, 28, 7243−7251
Energy & Fuels Article

concentrations gave a similar trend. This behavior could suggest


that the effect of the input O2 concentration on the SO3
production rate is of fractional order and not of first order with
respect to O2, as predicted by Burdett et al.17 Although Glueck
and Kenney37 proved independence of the SO3 production on
the O2 concentration, Svachula et al.34 and Forzatti35 stated
that a dependence upon O2 only manifests when the
concentrations of O2 and SO2 are comparable.
To evaluate the effect of the moisture content on the
conversion of SO2 to SO3, measurements between 3 and 9 vol
% were completed. Conversions at 0 vol % moisture (dry gas)
were not performed because of the need to condense the
sulfuric acid aerosols for the quantification of SO3 formed for
CCM. It can be seen in Figure 7 that increasing the moisture

Figure 5. Effect of the input SO2 concentration on the (a) output SO3
concentration and (b) conversion of SO2 to SO3. Inlet gas mixture for
the experiment: variable ppm of SO2, 5 vol % O2, 3 vol % H2O, and
balance N2, at a temperature of 900 °C.

Figure 7. Effect of the H2O concentration on the (a) output SO3


concentration and (b) conversion of SO2 to SO3. Inlet gas mixture for
the experiment: 1000 ppm of SO2, 5 vol % O2, variable vol % H2O,
and balance N2, at a temperature of 900 °C.

concentration from 3 to 9 vol % did not have much effect on


the conversion of SO2 to SO3. The effect of changing the H2O
concentration is not as pronounced as the effects of SO2 and O2
concentrations. Average conversions of 0.36−0.39% have been
observed. This suggests that the conversion of SO2 to SO3 is
independent of the water content being studied (between 3 and
9 vol %), although Forzatti et al.35 stated that the addition of
water inhibits SO2 conversion at low concentrations but levels
off at above 5% (v/v) water content. However, the absence of
the effect of the H2O concentration in the homogeneous
oxidation of SO2 was also evidenced in the works by Burdett et
Figure 6. Effect of the input O2 concentration on the (a) output SO3 al.,17 Schwaemmle et al.,33 and Svachula et al.34
concentration and (b) conversion of SO2 to SO3. Inlet gas mixture for 3.1.3. Effect of the Temperature: Conversions at 400−
the experiment: 1000 ppm of SO2, variable vol % O2, 3 vol % H2O,
and balance N2, at a temperature of 900 °C.
1000 °C. The temperature dependence of the homogeneous
oxidation of SO2 to SO3 was investigated at 1000 ppm of SO2, 5
vol % O2, 3 vol % H2O, and balance N2. The effect of the
input gives an increasing SO3 output. However, the SO3 output temperature is presented in Figure 8. The figure shows the SO2
data are similar to the kinetic prediction at lower O2 of 3% but to SO3 conversion data and a thermodynamic model, along
fall to 40% below the predicted value at an oxygen with the global kinetic model by Burdett et al.17 and its
concentration of 10%. The conversion at varying O2 input associated range of error (the upper and lower light dashed
7247 dx.doi.org/10.1021/ef5020346 | Energy Fuels 2014, 28, 7243−7251
Energy & Fuels Article

Figure 8. Homogeneous conversion of SO2 to SO3 at different temperatures. Inlet gas mixture for the experiment: 1000 ppm of SO2, 5 vol % O2, 3
vol % H2O, and balance N2. The Burdett et al.17 model was used for kinetics, and FactSage 6.3 software was used for the thermodynamic model.

Figure 9. Fly ash catalytic conversion of SO2 to SO3 at different temperatures. Fly ash used: A−air. Inlet gas mixture for the experiment: (air) 1000
ppm of SO2, 5 vol % O2, 3 vol % H2O, and balance N2 and (oxy) 1000 ppm of SO2, 5 vol % O2, 3 vol % H2O, 85 vol % CO2, and balance N2. Marier
and Dibbs settings: 1.8 s residence time, 185 cm3/min flow rate, and gas composition of 18.9 vol % H2O, 32.4 vol % O2, 8.4 vol % SO2, and 40.4 vol
% N2.16

lines). It can be seen that, thermodynamically, maximum SO3 is where catalysis by fly ash could enhance rapid formation of
expected at about 500 °C and, as the temperature is increased SO3.16,28
to 900 and 1000 °C, SO3 decreased to about 6.75 and 3.21%, In this regard, the effect of the temperature on the catalytic
respectively. conversion of SO2 to SO3 in the presence of fly ash supported
In summary, a positive correlation was found between SO3/ by quartz wool was investigated and is presented in Figure 9.
SO2 conversion and temperature. Increasing conversions from To simulate air and oxy-fuel flue gas compositions, the
0.04 to 0.77% were observed as the temperature was increased concentrations of the gas mixtures used in this study were
from 400 to 1000 °C during the post-flame homogeneous air
conversion experiments. Temperatures of around 900 °C or
greater are required to obtain an observable conversion with SO2 = 1000 ppm; O2
the available residence times provided by the uncatalyzed = 5 vol %; H 2O
homogeneous reactor. This is consistent with the findings by
Bayless et al.36 and Burdett et al.17 = 3 vol %; and balance N2
3.2. Heterogeneous Conversion (Fly Ash Effects).
oxy
Equilibrium calculations from 400 to 1000 °C in Figure 8
show that the maximum conversion of SO2 to SO3 happens at SO2 = 1000 ppm; O2
around 500 °C and that the equilibrium shifts to the formation
of SO2 at higher temperatures, approximately 10% at 900 °C.16 = 5 vol %; H 2O
While SO3 formation is favored at lower temperatures, the = 3 vol %; CO2
kinetics is relatively slow and said to be “frozen” at low
temperatures.6,16,38 Literature suggests that it is in this region, = 85 vol %; and balance N2

7248 dx.doi.org/10.1021/ef5020346 | Energy Fuels 2014, 28, 7243−7251


Energy & Fuels Article

The results of this study were plotted against a previous study Table 5. Homogeneous, Catalytic, and Total Conversions of
by Marier and Dibbs.16 Fly ash A derived from air firing is the SO2 to SO3 from 400 to 1000 °C Using Fly Ash A−Air
sample used for this plot. From Figure 9, it can be seen that, as
conversion (%)
with homogeneous conversion (Figure 8), the temperature has
a significant impact on the heterogeneous conversion of SO2 to temperature (°C) homogeneous total percent catalytic (%)
SO3 with fly ash as the catalyst. Marier and Dibbs16 found that 400 0.04 0.04 1.23
conversion is negligible at 500 °C and increases significantly at 500 0.06 0.08 19.21
700 °C with SO3/SO2 conversion of 27% before falling to 700 0.10 1.78 94.51
about 16% between 800 and 900 °C. Similarly, the post-flame 900 0.36 1.31 72.38
experiments in this study produced negligible SO3/SO2 1000 0.77 1.79 56.82
conversions at temperatures below 500 °C. A maximum
conversion of 1.7% at 700 °C was obtained before the total conversion (approximately 1.78%) of SO2 to SO3. With a
conversion decreased to 1.4% at 900 °C and then increasing further increase in the temperature, the catalytic component
again to about 1.6% at 1000 °C. It is however noteworthy that drops to 57% of the total conversion at 1000 °C.
the concentrations used by Marier and Dibbs16 were roughly 10 3.2.1. Effect of Fe2O3 on the Conversion of SO2 to SO3.
times greater than the concentrations typically found in the flue Iron oxide, Fe2O3, which is naturally occurring in fly ash of coal-
gas and that the residence time in their reactor was 1.8 s, as fired boilers is known to catalyze the formation of SO2 to
opposed to ∼1 s in this study; nonetheless, the temperature SO3.16 To observe the effect of the Fe2O3 content of the fly ash
dependence trend found in this study was similar. on the conversion of SO2 to SO3, fly ash samples A, B, and C
The effect of varying flue gas concentrations, i.e., gas from the three coals fired at different modes were investigated
concentrations equivalent to flue gas concentrations in air and and used in the study. Plotting the overall SO 3 /SO 2
oxy-fuel combustion, was also researched. Changing the gas conversions against the Fe2O3 content (Table 4) of the fly
composition from O2/N2 to O2/CO2 to simulate air and oxy- ash at 1000 ppm of SO2, 5 vol % O2, and 3 vol % H2O
fuel flue gas conditions did not have a significant effect on the conducted at 700 °C revealed that a general linear trend can be
conversion. Hence, for post-flame conditions, conversion of observed (Figure 11). A previous study16 using roughly 10
SO2 to SO3 is not a function of the mode of firing but a times higher gas concentrations showed a linear trend in their
function of the temperature and presence of the catalyst. It conversions using fly ash with a Fe2O3 content ranging between
could be observed from Figure 9 that the overall SO3/SO2 7 and 25 wt %.
conversions at 700 and 1000 °C are similar; however, it should
be noted that homogeneous and fly ash catalyses have
combined effects on the overall conversions.
Figure 10 presents the contribution of homogeneous and
catalytic effects on the overall conversions at different

Figure 11. Effect of the amount of Fe2O3 in fly ash on the catalytic


conversion of SO2 to SO3. Inlet gas mixture for the experiment: 1000
ppm of SO2, 5 vol % O2, 3 vol % H2O, and balance N2, at a
temperature of 700 °C. Marier and Dibbs settings: 1.8 s residence
time, 185 cm3/min flow rate, and gas composition of 18.9 vol % H2O,
32.4 vol % O2, 8.4 vol % SO2, and 40.4 vol % N2.16
Figure 10. Temperature effects on the catalytic and homogeneous
conversions of SO2 to SO3 using fly ash A−air. It can be observed that, even at lower Fe2O3 contents
typically found in the Australian coals and with 10 times lower
temperatures. The overall conversion at 400 °C started at gas concentrations, a linear trend was obtained (Figure 11).
0.04%, peaked at 1.78% at 700 °C, and then decreased to 1.31% The hollow triangle markers in the plot denote fly ash derived
at 900 °C, similar to the trend given by Marier and Dibbs.16 from air firing and the shaded triangle markers in the plot
This experiment tested the catalytic effect of fly ash A−air from denote fly ash obtained from oxy-fuel firing. It could be said
400 to 900 °C (Figure 9). However, extending the test to 1000 that, on the basis of the plot, the catalytic conversion of SO2 to
°C revealed an increase in average conversion to roughly 1.79%, SO3 is a function of the iron oxide content of the fly ash and
which is similar to the conversion obtained at 700 °C. It is that how the fly ash was derived, i.e., air-derived ash or oxy-fuel-
noteworthy that, although similar total conversions were derived ash, did not have any effects on the conversion.
observed at 700 and 1000 °C, the contribution of fly ash Nonetheless, apart from Fe2O3, which is widely known to have
catalysis is different. Table 5 shows that the conversion, which good catalytic effects, other metal oxides (e.g., CaO,13 TiO2,
could be attributed to fly ash, i.e., catalytic, is negligible at 400 and MnO) did not show any enhancement on the formation of
°C and greatest at 700 °C, accounting for about 95% of the SO3. With 0.5 g of fly ash for every 30 L of oxidant gases
7249 dx.doi.org/10.1021/ef5020346 | Energy Fuels 2014, 28, 7243−7251
Energy & Fuels Article

(STP), SO3 concentrations of 0.98 and 17.84 ppm were Carbon Capture and StorageCost Impacts and Coal Value; The
obtained at 700 °C for homogeneous and heterogeneous University of Newcastle: Callaghan, New South Wales, Australia, 2014.
conversions, respectively. This SO2 to SO3 conversion of 1.78% (2) Srivastava, R. K.; Miller, C. A.; Erickson, C.; Jambhekar, R. J. Air
falls within 0.5−3.5% conversion as stated by Spörl et al.10 and Waste Manage. Assoc. 2004, 54, 750.
(3) Hu, Y.; Naito, S.; Kobayashi, N.; Hasatani, M. Fuel 2000, 79,
between 1 and 3% from the 15 MWth pilot scale testing
1925.
conducted by Alstom.39 Therefore, with 0.5 g of fly ash (4) Reiner, T.; Arnold, F. J. Chem. Phys. 1994, 101, 7399.
containing 6.55 wt % Fe2O3 in the ash, a change in SO3 from (5) Zhuang, Y.; Pavlish, J. H. Environ. Sci. Technol. 2012, 46, 4657.
0.98 to 17.84 ppm at 3 vol % moisture and a contact time of 41 (6) Stanger, R.; Wall, T. Prog. Energy Combust. Sci. 2011, 37, 69.
ms with the ash is expected to change the ADP from 100 to 130 (7) Li, X.; Wall, T. F.; Stanger, R.; Liu, Y.; Belo, L.; Ting, T. S. S. Gas
°C, a 30 °C increase based on ZareNezhad’s correlation. Quality Control in Oxy-PF Technology for Carbon Capture and Storage;
The University of Newcastle, Callaghan, New South Wales, Australia,
4. CONCLUSION 2012.
(8) Belo, L. P.; Spörl, R.; Shah, K. V.; Elliott, L. K.; Stanger, R. J.;
The homogeneous (uncatalyzed) and heterogeneous (catalysis Maier, J.; Wall, T. F. Energy Fuels 2014, 28, 5472.
by fly ash) conversions of SO2 and O2 to SO3 in the absence of (9) Spörl, R.; Belo, L.; Shah, K. V.; Stanger, R.; Giniyatullin, R.;
combustibles were investigated experimentally between 400 Maier, J.; Wall, T. F.; Scheffknecht, G. Energy Fuels 2014, 28, 123.
and 1000 °C using a quartz flow reactor. (10) Spörl, R.; Walker, J.; Belo, L.; Shah, K.; Stanger, R.; Maier, J.;
Homogeneous (uncatalyzed) oxidation of SO2 to SO3 was Wall, T.; Scheffknecht, G. Energy Fuels 2014, 28, 5296.
examined and compared to a kinetic model by Burdett et al.17 (11) Fleig, D.; Alzueta, M. U.; Normann, F.; Abian, M.; Andersson,
The effect of increasing the SO2 concentrations gave an K.; Johnsson, F. Combust. Flame 2013, 160, 1142.
apparent order in SO 2 of greater than 1 at low SO 2 (12) Chem. Age (London) 1925, 13, 320.
(13) Jorgensen, T. L.; Livbjerg, H.; Glarborg, P. Chem. Eng. Sci. 2007,
concentrations, decreasing to a fractional order as the input
62, 4496.
concentration is increased. The O2 concentration was found to (14) Fleig, D.; Andersson, K.; Johnsson, F. Ind. Eng. Chem. Res. 2012,
have an influence on the oxidation of SO2 at concentrations of 51, 9483.
3−10 vol %. SO2 oxidation was also found to be independent of (15) Foster, P. M. Atmos. Environ. 1969, 3, 157.
the H2O concentration in the flue gas. Temperatures of at least (16) Marier, P.; Dibbs, H. P. Thermochim. Acta 1974, 8, 155.
900 °C are required for an observable conversion to occur for a (17) Burdett, N. A.; Langdon, W. E.; Squires, R. G. J. Inst. Energy
residence time of ∼1 s. 1984, 57, 373.
Heterogeneous (fly-ash-catalyzed) conversions were found to (18) Cullis, C. F.; Mulcahy, M. F. R. Combust. Flame 1972, 18, 225.
be greatest at 700 °C for the temperature range of 400−1000 (19) Rees, O. W.; Shimp, N. F.; Beeler, C. W.; Kuhn, J. K.;
°C, accounting for 95% of the overall conversions. Fly ash Helfinstine, R. J. In Circular 396; Frye, J. C., Ed.; Department of
coming from different modes of firing, i.e., air and oxy-fuel gas Registration and Education, State of Illinois: Springfield, IL, 1966.
(20) Spörl, R.; Maier, J.; Scheffknecht, G. Energy Procedia 2013, 37,
atmospheres, was found to have no effect on the conversion of
1435.
SO2 to SO3 when the concentration of SO2 was the same. (21) Krishnakumar, B.; Niksa, S. Proc. Combust. Inst. 2011, 33, 2779.
Similarly, the atmosphere associated with the mode of firing, (22) Ahn, J.; Okerlund, R.; Fry, A.; Eddings, E. G. Int. J. Greenhouse
i.e., predominantly N2 or CO2, had no impact on the Gas Control 2011, 5, S127.
conversion of SO2 to SO3. The presence of Fe2O3 catalytically (23) Wen, C. S.; Cowell, L. H. Sulphur Impact on Coal-Fired Gas
enhances conversion of SO2 to SO3. With greater SO2, H2O, Turbines; Solar Turbines Incorporated: San Diego, CA, 1987.
and ash loading during oxy-fuel, higher SO3 and, therefore, an (24) Flint, D.; Lindsay, A. W. Fuel 1951, 30, 288.
increased ADP are expected. (25) Goksöyr, H.; Ross, K. J. Inst. Fuel 1962, 35, 177.


(26) Fleig, D.; Andersson, K.; Normann, F.; Johnsson, F. Ind. Eng.
Chem. Res. 2011, 50, 8505.
AUTHOR INFORMATION
(27) Determination of sulfuric acid vapor or mist and sulfur dioxide
Corresponding Author emissions from kraft recovery furnaces; National Council of the Paper
*E-mail: terry.wall@newcastle.edu.au. Industry for Air and Stream Improvement, Inc. (NCASI). NCASI:
Research Triangle Park, NC, 1996.
Notes (28) Liu, Y.; Stanger, R.; Wall, T. Options for Sulphur Corrosion and
The authors declare no competing financial interest. Acid Dewpoint Monitoring in Oxy-fuel Combustion (Milestone Report);

■ ACKNOWLEDGMENTS
The authors acknowledge financial assistance provided by the
Australian National Low Emissions Coal Research and Development
(ANLEC R&D): Manuka, Australian Capital Territory, Australia, June
2011.
(29) Kommission Reinhaltung der Luft im VDI und DIN−
Xstrata Coal Low Emissions R&D Corporation Pty Ltd and Normenausschuss KRdL. Analysen- und Messverfahren II; Verein
through the Australian National Low Emissions Coal Research Deutscher Ingenieure (VDI)-Richtlinien: Düsseldorf, Germany, 2011.
and Development (ANLEC R&D). ANLEC R&D is supported (30) Determination of sulfuric acid and sulfur dioxide emissions from
by the Australian Coal Association Low Emissions Technology stationary sources; United States Environmental Protection Agency
Limited and the Australian Government through the Clean (U.S. EPA). U.S. EPA: Washington, D.C., p 733−753. (http://www.
Energy Initiative. The authors acknowledge the Institute of epa.gov/sam/pdfs/EPA-Method8.pdf).
Combustion and Power Plant Technology (IFK), University of (31) ZareNezhad, B. Oil Gas J. 2009, 107.
Stuttgart, Stuttgart, Germany, for the collaborative research (32) Suko, T.; Yamada, T.; Tamura, M.; Fujimori, T. Pilot Scale
initiative. Studies To Support Oxy-fuel Feasibility Study; Ishikawajima-Harima


Heavy Industries Co., Ltd.: Tokyo, Japan, 2006.
(33) Schwaemmle, T.; Heidel, B.; Brechtel, K.; Scheffknecht, G. Fuel
REFERENCES 2012, 101, 179.
(1) Shah, K.; Elliott, L. K.; Sporl, R.; Belo, L. P.; Wall, T. F. Coal (34) Svachula, J.; Alemany, L. J.; Ferlazzo, N.; Forzatti, P.; Tronconi,
Quality Impacts and Gas Quality Control in Oxy-fuel Technology for E.; Bregani, F. Ind. Eng. Chem. Res. 1993, 32, 826.

7250 dx.doi.org/10.1021/ef5020346 | Energy Fuels 2014, 28, 7243−7251


Energy & Fuels Article

(35) Forzatti, P. Catal. Today 1997, 34, 401.


(36) Bayless, D. J.; Jewmaidang, J.; Tanneer, S.; Birru, R. Proc.
Combust. Inst. 2000, 28, 2499.
(37) Glueck, A. R.; Kenney, C. N. Chem. Eng. Sci. 1968, 23, 1257.
(38) Cullis, C. F.; Henson, R. M.; Trimm, D. L. Proc. R. Soc. London,
Ser. A 1966, 295, 72.
(39) Levasseur, A. A.; Darling, S.; Kluger, F.; Moenckert, P.; Heinz,
G.; Prodhomme, B. Proceedings of the Coal-Gen Conference and
Exhibition; Columbus, OH, Aug 17−19, 2011.

7251 dx.doi.org/10.1021/ef5020346 | Energy Fuels 2014, 28, 7243−7251

You might also like