ArticlePDF Available

Shedding light on the governing mechanisms for insufficient CO and H2 burnout in the presence of potassium, chlorine and sulfur

Authors:

Abstract and Figures

Based on the experiences of insufficient burnout in industrial fluidized bed furnaces despite adequate mixing and availability of oxidizer, the influence of potassium on CO and H2 oxidation in combustion environments was investigated. The combustion environments were provided by a laminar flame burner in a range relevant to industrial furnaces, i.e. 845 °C to 1275 °C and excess air ratios ranging from 1.05 to 1.65. Potassium, in the form of KOH, was homogeneously introduced into the hot gas environments to investigate its effect on the radical pool. To quantitatively determine key species that are involved in the oxidation mechanism (CO, H2, KOH, OH radicals, K atoms), a combination of measurement systems was applied: micro-gas chromatography, broadband UV absorption spectroscopy and tunable diode laser absorption spectroscopy. The inhibition effect of potassium on CO and H2 oxidation in excess air was experimentally confirmed and attributed to the chain-terminating reaction between KOH, K atoms and OH radicals, which enhanced the OH radical consumption. The addition of chlorine or sulfur could reduce the concentrations of KOH and K atoms and consequently eliminated the inhibition on CO and H2 oxidation. Existing kinetic mechanisms underestimate the inhibiting effect of potassium and they fail to predict the effect of temperature on CO and H2 concentration when potassium and sulfur co-exist. This work advances the need to revise existing kinetic mechanisms to fully capture the interplay of K and S in the oxidation of CO and H2 in industrial fluidized bed furnaces.
Content may be subject to copyright.
Contents lists available at ScienceDirect
Fuel
journal homepage: www.elsevier.com/locate/fuel
Full Length Article
Shedding light on the governing mechanisms for insucient CO and H
2
burnout in the presence of potassium, chlorine and sulfur
Teresa Berdugo Vilches
a,
, Wubin Weng
b
, Peter Glarborg
c
, Zhongshan Li
b
, Henrik Thunman
a
,
Martin Seemann
a
a
Division of Energy Technology, Chalmers Tekniska Högskola, Sweden
b
Division of Combustion Physics, Lund University, P.O. Box 118, S-221 00 Lund, Sweden
c
Chemical Engineering, Technical University of Denmark, DK-2800 Kgs., Lyngby, Denmark
ARTICLE INFO
Keywords:
CO oxidation
Potassium
Inhibition
Combustion
UV absorption spectroscopy
TDLAS
ABSTRACT
Based on the experiences of insucient burnout in industrial uidized bed furnaces despite adequate mixing and
availability of oxidizer, the inuence of potassium on CO and H
2
oxidation in combustion environments was
investigated. The combustion environments were provided by a laminar ame burner in a range relevant to
industrial furnaces, i.e. 845 °C to 1275 °C and excess air ratios ranging from 1.05 to 1.65. Potassium, in the form
of KOH, was homogeneously introduced into the hot gas environments to investigate its eect on the radical
pool. To quantitatively determine key species that are involved in the oxidation mechanism (CO, H
2
, KOH, OH
radicals, K atoms), a combination of measurement systems was applied: micro-gas chromatography, broadband
UV absorption spectroscopy and tunable diode laser absorption spectroscopy. The inhibition eect of potassium
on CO and H
2
oxidation in excess air was experimentally conrmed and attributed to the chain-terminating
reaction between KOH, K atoms and OH radicals, which enhanced the OH radical consumption. The addition of
chlorine or sulfur could reduce the concentrations of KOH and K atoms and consequently eliminated the in-
hibition on CO and H
2
oxidation. Existing kinetic mechanisms underestimate the inhibiting eect of potassium
and they fail to predict the eect of temperature on CO and H
2
concentration when potassium and sulfur co-
exist. This work advances the need to revise existing kinetic mechanisms to fully capture the interplay of K and S
in the oxidation of CO and H
2
in industrial uidized bed furnaces.
1. Introduction
Combustion technologies are important sources of heat and elec-
tricity in todays society, and they are likely to play a role in the future
energy system as a stabilizing complement to intermittent energy
sources. To ensure that combustion takes place in the most ecient and
clean way, combustion units are subject of strict regulations on emis-
sions, e.g. limits on CO, NO
x
,SO
x
, dioxins. Among the regulated
emissions, CO should be kept below 150500 ppm at the stack for a
given excess air concentration of usually 6% for biomass and 11% for
waste. The exact values depend on the permissions for dierent plants
and are dened in national and local legislation, e.g. Regulation
(20013:253) for combustion of waste in Sweden [1].
A common reason for CO emissions is a local decit of O
2
due to
mixing limitations between fuel and air. Therefore, commercial com-
bustors operate with excess air, resulting in a concentration of some
percent O
2
at the stack. The problem of CO emissions is exacerbated in
facilities that combust biomass and inhomogeneous solid fuels such as
waste streams. In those cases, moderate temperatures (typically
750900 °C) are applied to reduce the risk of uncontrolled ash melt and
corrosion issues derived from the impurities in the fuel, as well as
prevent overheating of the furnace material; a measure that can also
challenge the complete oxidation of CO into CO
2
. Besides mixing and
temperature, CO oxidation can be inuenced by the chemical interac-
tion between the inorganic impurities in the fuel and the fuel oxidation
reactions. In fact, even trace levels of active species like alkali com-
pounds can drastically inuence the combustion chemistry in thermal
conversion processes [2].
The literature on the eect of alkali on the oxidation of CO is in-
conclusive and the underlying mechanisms are still under discussion.
Both, promoting and inhibiting eects of alkali species on CO oxidation
have been observed and are theoretically possible [2]. This reects the
strong dependence of the reactions on operating conditions, as well as
the complex interrelation between inorganic and organic chemistry.
https://doi.org/10.1016/j.fuel.2020.117762
Received 30 January 2020; Accepted 31 March 2020
Corresponding author.
E-mail address: berdugo@chalmers.se (T. Berdugo Vilches).
Fuel 273 (2020) 117762
0016-2361/ © 2020 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/BY/4.0/).
T
Previous work in the 12-MW Chalmers circulating uidized bed (CFB)
boiler [3,4] has evidenced the strong inhibiting eect of K on CO oxi-
dation during combustion of biomass in a bed of olivine. Potassium was
introduced in the combustor as part of the fuel ash and it accumulated
in the bed material, wherefrom it could be released in gas-phase to the
reaction environment [4].Fig. 1 shows the increasing levels of CO over
time, as the ash elements accumulate in the olivine bed. Despite the
temperature (850 °C) and oxygen concentrations (3.5% vol) were the-
oretically sucient to ensure a complete oxidation of the fuel, CO
concentrations steadily increased reaching values above 1000 ppm at
the stack [3]. The causality between the CO emissions and the presence
of K was further conrmed by adding small batches of K salts to the
uidized bed (not shown in the gure). Similar inhibitory eect of CO
by Na salts have been observed in smaller units, e.g. in a laboratory-size
uidized bed reactor by Bulewiz et al. [5], and in a 300-kW down-red
radiant furnace by Lissianski et al. [6]. Hindiyarti et al. [7] found that
the inhibiting eect of K is also relevant to the oxidation of CO by water
vapour in the absence of oxygen in the temperature range 5001100 °C.
The inhibiting eect was proportional to the concentration of po-
tassium and it levelled oat high concentrations (above 500 ppm under
the investigated conditions). Contrarily, Ekvall et al. [8] did not observe
an inuence of K on the oxidation of CO during air-combustion of
propane in a 100-kW experimental rig at temperatures above 1000 °C,
while a promoting eect was found under oxy-fuel conditions.
The inhibiting eect of alkali is mostly attributed to gas-phase ra-
dicals removal reactions [2,9,10]. According to Glarborg [2], the main
radical removing reactions are R1 and R2. The sequence corresponds to
the overall step H + OH H
2
O, where two radicals are removed.
KOH+HK+H
2
O (R1)
K+OH+MKOH+M (R2)
H+O
2
OH+O (R3)
CO+OHCO
2
+H (R4)
Other trace species than alkali, e.g. sulfur and chlorine, are also
known to inuence the CO oxidation [2]. The experience of the Chal-
mers uidized bed boiler shown in Fig. 1 exemplies the strong re-
sponse of the CO emissions to the addition of elemental sulfur. With
1 kg/h of sulfur the CO emissions went rapidly from 1200 ppm to zero,
in a system containing 3 tons of olivine bed and fed with 1.8 ton/h of
biomass. A similar drastic response was observed in a 26-kW bench-
scale diusion ame during combustion of natural gas under slight sub-
stochiometric conditions and with the addition of 0100 ppm of SO
2
[11]. In contrast, other work has reported that the addition of SO
2
or
higher levels of sulfur in the fuel inhibit CO oxidation, in a laboratory
ow reactor [12] and in the uidized bed reactors [13,14], respec-
tively. Alzueta et al. [15] describe that sulfur promotes the oxidation of
CO under nearly stochiometric conditions, while it mildly inhibits it
under fuel lean and rich conditions. Yet, this explanation disagrees with
the observations at the Chalmers boiler, where the promoting eect of S
on CO oxidation occurred with excess air (3.5% vol in the ue gases or
excess air ratio 1.2).
Most previous research indicates that the sulfur chemistry has a
direct inuence on CO oxidation [11,15], while the experience at
Chalmers with K-containing biomass and addition of S [3] also points at
interrelations between the S and K chemistry that aect the CO oxi-
dation. Hypothetically, this can occur via e.g. sulfation reactions [8,16]
that alter the speciation of K in the reaction environment. Similarly,
other species present in biomass such as chlorine could also interfere
with the K chemistry [12] and, thereby, inuence the CO oxidation.
Chlorine by itself have been found to inhibit CO oxidation in a number
of investigations. Under moist conditions, Roesler et al. [17] found that
trace levels of HCl (0200 ppm) at atmospheric pressure clearly in-
hibited CO oxidation in the temperature range of 500900 °C. Wu et al.
[18] found that the inhibiting eect was more pronounced in oxygen-
rich conditions than at stochiometric or oxygen-lean conditions. The
inhibiting eect of chlorine has also been observed under uidized bed
operation in the CANMET 0.8 MW CFB combustor using coal as fuel and
operating at 850 °C [19,20].
Biomass and waste fuels typically contain chlorine, sulfur and po-
tassium. An understanding of the CO oxidation in large scale combus-
tors calls for a better knowledge of the interplay of these species on the
oxidation chemistry. The aim of this work is to shed light on the gov-
erning mechanisms for insucient CO burnout due to the presence of K,
despite adequate mixing and availability of oxidizer that has been ob-
served in large scale combustors. A laminar burner system was em-
ployed to map the inuence of K in the oxidation of CO covering re-
action conditions relevant to combustion of biomass and solid waste in
uidized bed combustors, i.e., mild temperature, excess air and co-ex-
istence of trace elements. CO oxidation was investigated in the presence
of K, S and Cl. The work involves a quantitative study in combustion
environments with well-controlled temperature and oxidation condi-
tions, with quantitative measurement of KOH, K-atom, OH radical, CO
and H
2
. The suitability of existing kinetic mechanisms to explain the
results was assessed and discussed.
2. Experimental
Experiments were carried out with a CH
4
/N
2
/O
2
ame generated in
a multi-jet burner and the ue gas composition was analysed under
various ame conditions, notably temperature and excess air (λ). The
conditions investigated are summarized in Table 1, where the excess air
ratio ranges from 1.04 to 1.67 and the temperature ranges from 1275 to
845 °C. The range was chosen to cover λ= 1.2 and temperature of
845 °C, as they correspond to the ue gas conditions typically found in
commercial uidized bed combustors, as well as those applied in the
Chalmers system when the eect of K and S has been evidenced.
Seeding of K-, S- and Cl-containing species were conducted under
the ame conditions described in Table 1 to investigate their eect on
the oxidation of CO. Further details on the seeding system are described
in subsection 2.2. Two sets of experiments were dened to system-
atically address two questions: (1) the inuence of potassium on the
oxidation of CO at dierent ame conditions; and (2) the inuence of
sulfur and chlorine, respectively, on the oxidation of CO in co-existence
Fig. 1. Inuence of accumulation of K in the bed material on the CO emissions
from the 12-MW Chalmers circulating uid bed boiler, and eect of elemental S
in mitigating the CO emissions. Combustion of woody biomass at 850 °C with
excess air, i.e. 3.5%vol O
2
at the stack. Bed material: Olivine [3].
T. Berdugo Vilches, et al. Fuel 273 (2020) 117762
2
with potassium. The series of experiments carried out in each set are
listed in Table 2. In the rst set, potassium was rstly added in dierent
concentrations at constant temperature and excess air ratio (series K1);
secondly, the excess air ratio was varied at constant temperature and
with a xed seeding rate of potassium (series K2); and thirdly, the
temperature was varied at constant excess air ratio and seeding rate of
potassium (series K3). In the second experimental set, the inuence of
sulfur in co-existence with a xed seeding ow of potassium was in-
vestigated at dierent temperatures while keeping the excess air ratio
constant (series SK). Finally, the eect of chlorine was investigated with
a single test (labelled ClK in Table 2) that involved the simultaneous
addition of K and Cl, and that was compared to a similar case with
addition of K and in the absence of Cl.
The concentrations of CO and H
2
in the ue gas was measured with
a micro gas chromatograph (µ-GC); the concentration of K atoms by a
TDLAS system; and the concentrations of KOH, KCl and OH radicals by
UV absorption spectroscopy. The burner, seeding system and the
measurement techniques applied are described in detail below.
2.1. Burner
The multi-jet burner applied enables the generation of a hot gas
environments with an even temperature distribution. The detailed
structure of the burner was described by Weng et al. [21], and only a
brief description is given here. The burner was comprised of two
chambers, namely the jet chamber and the co-ow chamber. The pre-
mixed CH
4
/air/O
2
gas mixtures were introduced into the jet chamber
and evenly distributed to 181 steel tubes with an inner diameter of
1.6 mm. Premixed ames anchored on the jet tubes to provided hot ue
gas, which generated a hot gas environment above the outlet
(85 mm × 47 mm) of the burner for the present study. Each tube was
surrounded by even co-ow gases, air/N
2
, which was introduced
through the co-ow chamber. All the gas ows were controlled by mass
ow controllers (Bronkhorst). The temperature information was ob-
tained through a two-line atomic uorescence thermometry system of
indium atoms as described by Borggren et al. [22].
2.2. Seeding of K, S and Cl
To introduce potassium into the hot gases, a water solution of po-
tassium carbonate with a concentration of 0.5 mol/l was used. As
shown in Fig. 2, solution fogs were generated through an ultrasonic
fogger and transported by air into the jet chamber. As the potassium
carbonate passed through the ames, gas-phase potassium hydroxide
can be formed in the hot gas. The potassium concentration in the hot
gas could be adjusted from 0 to 31 ppm by varying the ow rate of the
air for transporting solution fogs. The concentration of total potassium
seeded into the hot ue gas was estimated from the measurements of
KOH and K atoms. The seeding of trace amount of K
2
CO
3
was observed
to have negligible eect on the ue gas temperature.
In the experiments with sulfur, SO
2
was seeded from a gas cylinder
into the hot gas after the mixing with co-ow air/N
2
. To seed chlorine
to the ame, chloroform (CHCl
3
) was added into the hot gas system.
The CHCl
3
liquid was stored in a bubbler and kept at 10 °C using a
chilled bath (PolySicence). A N
2
ow was used to carry CHCl
3
vapor
into the jet chamber. Potassium chloride was generated by the co-
seeding of potassium carbonate. The seeding of sulfur was done at a
concentration of 150 ppm while the seeding of chlorine was approxi-
mately 400 ppm, which could fully convert KOH in to K
2
SO
4
or KCl at
the temperature of 845 °C.
2.3. GC system
The micro-gas chromatograph (µ-GC) was a Varian CP4900 with
two channels equipped with a PoraplotQ and a M S5Å column, and
using He and Ar as carrier gases, respectively. A sampling tube made by
stainless steel was used to extract the ue gas at a height of about 5 mm
above the burner outlet. A gas sample of approximately 300 mL was
collected in a gas bag with the help of a syringe. During sampling, the
gas was cooled down in the sampling tube, and it was ltered and dried
in a LC-NH2 adsorption column before entering the gas bag. Thereafter,
the gas bag was connected to the inlet of the µ-GC, so the instrument
could take samples from it.
For each test, one gas bag was lled with a sample gas, and its
content was analysed ve times, i.e. ve repeat chromatograms were
created per ame conditions. Each chromatogram is the result of a
point-injection (1030 ms sampling time), which was taken in intervals
of 4 min to allow the necessary analysis time required by the µ-GC. The
concentrations of H
2
and CO presented in this work for each test are the
average of the ve repeat chromatograms, and the error bars represent
the corresponding standard deviation.
2.4. TDLAS system
The TDLAS system applied in the measurement of K atoms has been
previously described in refs. [23,24]. In this system, a laser at wave-
length of 769.89 nm was used to detect the potassium atoms with its
transition of 4
2
S
1/2
4
2
P
1/2
. The laser was provided by an external
cavity laser (Toptica, DL100), and it had a power of about 3 mW and a
beam size of about 1 mm
2
. The cavity laser was controlled by an
Table 1
Flame conditions investigated.
Flame Case Gas ow rate (sl/min) Excess
air ratio
λ
Gas product
Temperature T
(°C)
Jet-ow Co-ow
CH
4
Air O
2
N
2
Air
T1O1 2.69 13.29 2.80 30.30 0.00 1.04 1275
T2O1 2.48 12.91 2.45 35.35 0.00 1.04 1115
T3O1 2.27 11.84 2.25 40.40 0.00 1.04 985
T4O1 1.86 9.68 1.83 40.40 0.00 1.04 845
T4O2 1.86 9.68 1.83 35.35 5.04 1.32 845
T4O3 1.86 9.68 1.83 29.24 11.13 1.67 845
Table 2
Experimental series according to the question investigated.
Experimental sets Series Variable Range of variation Flame conditions* Others
(1) Eect of potassium on CO oxidation K1 Potassium seeding 032 ppm T4O1
K2 Excess air ratio 1.041.67 T4O1, T4O2, T4O3 K-seeding11 ppm
K3 Temperature 8451275 °C T1O1, T2O1, T3O1, T4O1 K-seeding 11 ppm
(2) Eect of Cl and S on CO oxidation in co-existence with K SK Temperature 8451275 °C T1O1, T2O1, T3O1, T4O1 K-seeding 11 ppm
S-seeding 150 ppm
ClK ––T4O1 K-seeding 11 ppm
Cl-seeding 400 ppm
*Flame conditions described in detail in Table 1.
T. Berdugo Vilches, et al. Fuel 273 (2020) 117762
3
analogy control package with a temperature control module (DTC 110),
a current control module (DCC 110), and a scan module (SC 110). The
scan module was adopted to make the laser have scanning of the wa-
velength with a range over 35 GHz at a repetition rate of 100 Hz. The
power of the laser was monitored by two photodiodes (PDA100A,
Thorlabs), and the scanning range was measured by a high-nesse
confocal Fabry-Perot etalon (Topoca, FPI 100), shown in Fig. 2(b). As
the laser passed through the hot ue gas containing potassium atoms,
the light was absorbed and the concentration of potassium atoms can be
derived through the Beer-Lambert law [24,25]. In the present study, the
laser beam located at about 5 mm above the burner outlet with a path
length of 8.5 cm. The potassium line was determined to have a Voigt
prole with a full width at half maximum (FWHM) of 5.34 GHz.
2.5. UV absorption spectroscopy system
In the experiments, the concentrations of K atoms, KOH, KCl and OH
radicals were measured by UV absorption spectroscopy in the hot gas
environments. As shown in Fig. 2(c), a deuterium lamp was used to
generate a collimated UV light with a beam size of about 5 mm. The UV
light was guided by ve UV-enhanced aluminium mirrors to pass
through the hot gas with a path length of 522 mm [26]. After the
passage, the light was collected and analysed by a spectrometer.
The experimental absorbance data were obtained through the di-
vision between the intensity of UV light after the passage of the ame
with potassium seeding and the one without potassium seeding. It was
used to determine the concentration of KOH and KCl in the hot gas
through the Beer-Lambert law. In the calculation, the UV absorption
cross-sections of KOH and KCl measured by Weng et al. [24] was used.
Typical UV absorption spectrum obtained in KOH concentration mea-
surement are shown in Fig. 3 represented by the black dots. Two ab-
sorption peaks centred at around 250 nm and 330 nm can be observed
due to the absorption of gas-phase KOH, as described in our previous
studies. The absorbance of KOH was simulated using its UV absorption
cross-sections, and it could be well overlapped to the measured results
as the concentration was set to 11.5 ppm (see Fig. 3 with a red line).
The dierences between the measured one and the simulated one are
the additional absorption at around 404 nm and two dips at around
283 nm and 310 nm. The former one was from the K-atom absorption,
while those two dips were formed due to the reduction of OH radicals in
the ame after the seeding of potassium. In the measurement of OH
radicals, the absorption at around 310 nm were obtained through the
division between the intensity of UV light after the passage of the ame
and the one without ame. A typical experimental result was shown in
the inset of Fig. 3 with black dots. This measured OH absorbance can be
tted (as shown by the solid line) to retrieve the OH concentration
Fig. 2. Schematic of the experimental setup. (a) The structure of the burner system with GC sampling system. (b) TDLAS system for K-atom measurement. (c) UV
absorption spectroscopy system for quantitative measurements of KOH, KCl and OH radical.
Fig. 3. UV absorption spectrum obtained for KOH and OH measurement. The
measured absorbance values were presented with black dot and the corre-
sponding tting curves based on UV-absorption cross-section of KOH [24] and
OH [27,28] were shown with red lines. (For interpretation of the references to
colour in this gure legend, the reader is referred to the web version of this
article.)
T. Berdugo Vilches, et al. Fuel 273 (2020) 117762
4
through LIFBASE [28]. As a result of the line-of-sight nature of the
absorption spectroscopy technique, the concentration obtained by this
method is the average concentration along the horizontal direction
from the centre to the edge of the hot ue gas.
3. Modelling
The simulations were conducted using Chemkin PRO [29] as de-
scribed by Weng et al. [30]. A one-dimensional stagnation reactor in
Chemkin PRO was adopted to simulate the reaction occurring in the hot
ue gas along the central axial direction. The gas inlet of the reactor
was assigned to be the mixture of the hot ue gas at 3 mm from the jet
ame fronts and the co-ow gas. The composition of the hot ue gas
produced from the premixed jet ames was obtained using a one-di-
mensional free propagation premixed ame model in Chemkin PRO
with the GRI-3.0 mechanism [31]. The temperature along the axial
direction in the reactor was set to be the one measured by a type B
thermocouple in the burner. The measured temperature was corrected
based on the heat transfer theory due to heat losses by thermal radia-
tion. The axial distance in the reactor was set to 64 mm, which was the
distance between the ame front and the ow stabilizer. The horizontal
distribution of the species concentration was obtained using a one-di-
mensional opposed-ow model in Chemkin PRO as reported by Weng
et al. [30]. Through this model, the reaction between the hot ue gas
and the ambient air on the edge of the hot ue gas was mimicked.
Average concentrations on the horizontal direction were derived from
the horizontal distribution of species obtained in the simulations to
obtain a result that could be compared to the measurements by ab-
sorption spectroscopy. Thus, both the theoretical and experimental
concentrations shown in the results section refer to the average on the
horizontal direction.
The modelling was conducted with a chemical kinetic model that
accounted for both the KClS interactions, including chlorination and
sulfation of potassium, and the impact of the trace species on the O/H
radical pool. With a few changes, the KClS mechanism reported by
Weng et al. [16] was chosen for the calculations. This mechanism draws
on previous work on alkali metals [3234] as well as sulfur [35] and
chlorine [36] chemistry. In the present work, the alkali subset was re-
vised, with a particular emphasis on the interaction of potassium spe-
cies with the radical pool. The full mechanism is available as
Supplementary Material.
Reactions of potassium species with the radical pool were discussed
in previous work [2,7,32]. Few of these steps have been characterized
experimentally, and most rate constants have been estimated by ana-
logy with reactions of sodium and lithium. Most of the K/O/H subset is
the same as that proposed by Glarborg and Marshall [32]. However, the
thermochemistry for KO
2
and the rate constant for
K+O
2
+M=KO
2
+ M were drawn from the more recent work of
Sorvajärvi et al. [37]. Under oxidizing conditions, the KO
2
radical may
play an important role due to its relatively high thermal stability. Alkali
superoxides have been reported to enhance radical removal in pre-
mixed ames [3841] and KO
2
is believed to participate in the reaction
sequences leading to sulfation of KOH [42]. While the K + O
2
+M
reaction is now well established, there are no measurements available
for the chain terminating step KO
2
+OH=KOH+O
2
. In the present
work, we have chosen a rate constant for KO
2
+OHof10
14
cm
3
mol
1
s
1
; this value is possibly an upper limit. Another reaction of interest is
the oxidation of CO by KO
2
to form CO
2
+ KO. Perry and Miller [43] set
this step to be very fast, with a rate constant of 10
14
cm
3
mol
1
s
1
, but
their modelling predictions were not sensitive to the value. The Perry
and Miller rate constant, which is suciently fast to eliminate inhibi-
tion of CO oxidation by potassium species under the present conditions,
is not supported by any experimental or theoretical work, and we be-
lieve that a more realistic value is 2 × 10
12
cm
3
mol
1
s
1
.
Fig. 4. The variation of the concentration of K atom and OH radical with dif-
ferent amount of potassium seeded into the hot gas at temperature of 845 °C
and excess air ratio of 1.04.
Fig. 5. The variation of the concentration of CO and H
2
with dierent amount
of total potassium seeded into the hot gas at temperature of 845 °C and excess
air ratio of 1.04.
Fig. 6. The variation of the concentration of K atoms and OH radical at dif-
ferent excess air ratio with a temperature of 845 °C and 11 ppm K seeding.
T. Berdugo Vilches, et al. Fuel 273 (2020) 117762
5
4. Results
The concentration of CO, H
2
, KOH, OH and K for the experimental
series listed in Table 2 are shown in Figs. 411. The measurements are
presented and discussed in two parts corresponding to the two sets of
experiments in Table 2: those related to the inuence of potassium on
CO oxidation at dierent operating conditions; and those related to the
inuence of sulfur and chlorine in co-existence with K.
4.1. The eect of operating conditions on inhibition by KOH
Fig. 4 shows results for K atoms and OH measured in a hot ue gas
temperature of 845 °C and an excess air ratio of 1.04, i.e. series labelled
as K1 in Table 2. The concentration of K atoms increased with the total
potassium seeded but levelled out at high potassium loadings. The
concentration of OH decreased strongly as the total potassium con-
centration was increased from 0 to 2 ppm. With potassium levels above
2 ppm, the OH radical concentration in the hot gas was below the de-
tection limit. This shows that the presence of K atoms strongly enhances
the consumption of OH radicals in the hot gas environment.
The corresponding concentrations of CO and H
2
, measured by the µ-
GC, are shown in Fig. 5. The levels of CO and H
2
increased from 0 ppm
to 1800 ppm and 700 ppm, respectively, as the potassium seeding in-
creased to 31 ppm. As the concentration of potassium in the ame in-
creases, the concentrations of CO and H
2
increase until they level o.
This trend is in agreement with that obtained by Hindiyarti et al. [7] in
Fig. 7. The variation of the concentration of CO and H
2
at dierent excess air
ratio with a temperature of 845 °C and 11 ppm K seeding.
Fig. 8. The variation of the concentration of KOH (a) and K atoms (b) at dierent temperature and SO
2
seeding at excess air ratio of 1.04 and about 11 ppm of K
seeding.
Fig. 9. The variation of the concentration of OH radical at dierent tempera-
ture with an excess air ratio of 1.04. The red line indicates the case without K
and SO
2
seeding; black line indicates the case with K seeding but without SO
2
seeding; blue line indicates the case with K and SO
2
seeding. (For interpretation
of the references to colour in this gure legend, the reader is referred to the web
version of this article.)
Fig. 10. The variation of the concentration of CO at dierent temperature with
an excess air ratio of 1.04. The red line indicates the case without K and SO
2
seeding; black line indicates the case with K seeding but without SO
2
seeding;
blue line indicates the case with K and SO
2
seeding. (For interpretation of the
references to colour in this gure legend, the reader is referred to the web
version of this article.)
T. Berdugo Vilches, et al. Fuel 273 (2020) 117762
6
moist conditions and in the absence of oxygen. In their case, the con-
centration of potassium required to saturate the inhibiting eect of
potassium was one order of magnitude larger than that obtained in this
work in the presence of oxygen. Note that the increase in the levels of
unburned CO and H
2
is consistent with the depletion of OH shown in
Fig. 4. The results conrm that potassium has a strong inhibiting eect
on the oxidation of CO and H
2
.
The modelling predictions are in qualitative agreement with the
measurements, but there are quantitative dierences. While the K atom
concentration is slightly underpredicted, the impact of K on the OH
level is signicantly smaller than observed. The simulations correctly
predict a signicant inhibition of the oxidation of CO by the addition of
potassium, but the eect on H
2
is underpredicted.
The inhibiting eect of potassium on the oxidation of CO/H
2
in the
hot ue gas was also observed at higher excess air ratios, i.e.,
λ= 1.041.67, as shown in Fig. 6. In this experimental series (series K2
in Table 2), the excess air was adjusted by modifying the amount of air
and N
2
in the co-ow. The temperature was kept constant to 845 °C and
the potassium seeding to 11 ppm. The concentration of K atoms de-
creased signicantly with the increase of excess air ratio. However, all
the cases showed a signicant consumption of OH. The concentrations
of CO and H
2
are shown in Fig. 7, where the inhibiting eect becomes
weaker with the increase of excess air ratio. Under the conditions
tested, the inhibiting eect of K on the oxidation of H
2
can no longer be
detected at excess air ratios above 1.35.
The simulation results are in reasonable agreement with the ex-
perimental data. Similar to the observations at λ= 1.04, the OH
concentration in the presence of potassium is overpredicted, while at
zero KOH feeding, OH is predicted within 30%.
In the experimental series K3 (see Table 2), the temperature was
varied in the range 8451275 °C, while the excess air ratio was kept at
1.04 and the seeding of K
2
CO
3
corresponded to 11 ppm of total po-
tassium. The resulting concentrations of KOH and K atoms are pre-
sented in Fig. 8, and the concentration of OH in the cases with and
without K is shown in Fig. 9 (see cases without sulfur seeding in the
gures). Potassium can enhance the consumption of OH under dierent
temperatures. The corresponding concentrations of CO and H
2
in the
hot gas with and without potassium seeding are shown in Fig. 10 and
Fig. 11, respectively. The inhibition eect was weakened with the in-
crease of temperature and there is almost no eect as the gas tem-
perature increased to 1275 °C. This is in line with the ndings of Ekvall
et al. [8] that found no inhibiting eect by potassium at temperatures
above 1000 °C.
The model predicts the variation in the concentration of potassium
and OH radicals as a function of ue gas temperature satisfactorily. The
dierence between the trends of the concentration of K atoms obtained
in the simulation and in experiment might be caused by the small in-
homogeneity of the oxygen concentration in the hot ue gas con-
sidering the strong inuence of oxygen concentration on the K atom
concentration as shown in Fig. 6, but still the simulation and experi-
mental results were at the same concentration level and much smaller
than the concentration of KOH as shown in Fig. 8.
4.2. Inuence of sulfur and chlorine on K-inhibition
The results derived from the tests with seeding of sulfur are shown
in Figs. 811, where they are labelled as with SO
2
. In the cases with co-
seeding of potassium (series SK in Table 2), the addition of sulfur show
larger inuence in the low temperature range compared to the cases at
higher temperature. For instance, with the addition of sulfur at 845 °C
the concentrations became closer to the conditions without potassium
seeding (compare black and blue solid lines in Figs. 1011). This ob-
servation is in line with the sulfation reaction [8]. At 845 °C, the pre-
sence of sulfur reduces the formation of K atoms and promotes con-
version of KOH to K
2
SO
4
particles. Consequently, more OH survived at
this temperature, as shown in Fig. 9. Due to the removal of gas phase
potassium by the sulfation process the inhibitory eect of potassium is
mitigated.
The results prove the interplay between potassium and sulfur in the
CO oxidation under the conditions tested, and conrm the hypothesized
mechanism involved in the large scale tests at the Chalmers unit [3].
Therefore, the direct eect of sulfur in the oxidation process, as de-
scribed by several authors for fuels that are free of impurities [1115] is
not sucient to capture the inuence of sulfur in combustion units with
more complex fuels. The complexity of the reaction process is increased
by the dependence of the behaviour of sulfur at dierent temperatures
as seen in Figs. 1011 and air excess ratio as described by Alzueta et al.
[15].
In the test with seeding of chlorine (ClK in Table 2), the presence of
approximately 400 ppm of Cl caused all the potassium in the hot gas to
be converted into KCl, which was measured by the UV absorption
system, as shown in Fig. 8. After seeding of Cl, a much smaller level of K
atoms was detected; about 30 times smaller compared to the KOH case.
At the same time, the OH concentration increased from 0 ppm to
10 ppm, as presented in Figs. 8 and 9, while the H
2
and CO con-
centrations reduced to 0 ppm (cf. Figs. 10 and 11). This is in contrast to
the inhibiting eect of Cl on the oxidation of CO commonly reported in
literature [1720], which emphasizes the relevance of the combined
eect of active species in complex systems. The data show clearly that
the presence of sulphur or chlorine eectively eliminate the inhibition
eect of potassium on the CO/H
2
.
The model predicts correctly the removal of KOH and K atoms with
SO
2
or Cl seeding, especially at the temperature of 845 °C. Potassium
sulfate and potassium chlorine are more stable in the gas phase than
potassium hydroxide, resulting in lower predicted concentrations of K
atoms in the hot ue gas and higher levels of OH radicals (cf. Fig. 9). As
a result, the inhibition eect on CO oxidation is weakened (cf. Fig. 10).
However, at the temperature above 985 °C, a larger deviation between
the simulation and experimental results was observed.
5. Conclusion
This work sheds light on the true mechanism behind insucient
combustion in industrial furnaces in the presence of potassium despite
adequate mixing and availability of oxidizer. It is demonstrated that
trace level of potassium, i.e. tenths of ppm, inhibits the gas-phase oxi-
dation of CO and H
2
in the temperature range of 8451275 °C and
equivalence ratio of 1.041.65. Up to 2000 ppm CO and 700 ppm H
2
was observed to survive the oxidation in oxygen rich hot environments
Fig. 11. The variation of the concentration of H
2
at dierent temperature with
an excess air ratio of 1.04. The red line indicates the case without K and SO
2
seeding; black line indicates the case with K seeding but without SO
2
seeding;
blue line indicates the case with K and SO
2
seeding. (For interpretation of the
references to colour in this gure legend, the reader is referred to the web
version of this article.)
T. Berdugo Vilches, et al. Fuel 273 (2020) 117762
7
due to the co-existence of trace amount of potassium. The experiments
showed that higher temperature and higher oxygen concentration, re-
spectively, weakens the inhibiting eect of potassium. The reduction of
OH concentration due to the existence of K atoms is experimentally
measured and identied as the main cause for the inhibition of the
oxidation of CO and H
2
. It was further shown that sulfation of the po-
tassium species led the eect to be eliminated. These ndings provide
an explanation to the casualty between increasing CO concentration
and the presence of potassium in ue gases, as well as to the positive
impact of sulfur additions in large scale combustion units on the CO
emissions. Existing kinetic models capture the qualitative interrelation
between potassium, OH radicals, CO and H
2
, respectively, at 845 °C and
excess air ratio 1.04, as well as the eect of varying excess air ratio.
However, the inhibiting eect of potassium is more intense in practice
than what the predictions indicate. The model fails to predict the ex-
perimental trends at varying temperature when sulfur and potassium
coexist in the combustion environment, and the predictions of H
2
concentration are generally less accurate than those for CO. Overall,
this work serves as a basis for the improvement of the theoretical de-
scription of the reaction mechanism involved in the inhibition of H
2
and
CO oxidation by potassium species and calls for the revision of existing
mechanisms.
Declaration of Competing Interest
The authors declare that they have no known competing nancial
interests or personal relationships that could have appeared to inu-
ence the work reported in this paper.
Acknowledgements
The research leading to these results received funding from the
Swedish Energy Agency (STEM) through the CECOST project and the
Swedish gasication centre SFC.
Appendix A. Supplementary data
Supplementary data to this article can be found online at https://
doi.org/10.1016/j.fuel.2020.117762.
References
[1] M.-o. energidepartementet, Förordning (2013:253) om förbränning av avfall, in: M.-
o. energidepartementet (Ed.), 2013.
[2] Glarborg P. Hidden interactionstrace species governing combustion and emis-
sions. Proc Combust Inst 2007;31:7798.
[3] Thunman H, Seemann M, Berdugo Vilches T, Maric J, Pallares D, Ström H, et al.
Advanced biofuel production via gasication lessons learned from 200 man-years
of research activity with Chalmersresearch gasier and the GoBiGas demonstration
plant. Energy Sci Eng 2018;6:634.
[4] Marinkovic J, Thunman H, Knutsson P, Seemann M. Characteristics of olivine as a
bed material in an indirect biomass gasier. Chem Eng J 2015;279:55566.
[5] Bulewicz EM, Janicka E, Kandefer S. Halogen inhibition of CO oxidation during the
combustion of coal in uidized bed; 1989. p. 1638.
[6] Lissianski VV, Zamansky VM, Maly PM. Eect of metal-containing additives on nox
reduction in combustion and reburning. Combust Flame 2001;125:111827.
[7] Hindiyarti L, Frandsen F, Livbjerg H, Glarborg P. Inuence of potassium chloride on
moist CO oxidation under reducing conditions: experimental and kinetic modeling
study. Fuel 2006;85:97888.
[8] Ekvall T, Andersson K. Experimentally observed inuences of KCl and SO2 on CO
oxidation in an 80 kW oxy-propane ame. In: International Pittsburgh Coal
Conference, Pittsburgh, USA; 2017.
[9] Friedman R, Levy JB. Inhibition of opposed jet methane-air diusion ames. The
eects of alkali metal vapours and organic halides. Combust Flame
1963;7:195201.
[10] McHale ET. Flame inhibition by potassium compounds. Combust Flame
1975;24:2779.
[11] Andersen J. Emission control through sulfur addition. M.Sc. Thesis, Department of
Chemical Engineering, Technical University of Denmark, 2006.
[12] Dam-Johansen K, mand L-E, Leckner B. Inuence of SO2 on the NON2O chemistry
in uidized bed combustion: 2. Interpretation of full-scale observations based on
laboratory experiments. Fuel 1993;72:56571.
[13] Anthony EJ, Lu Y. Relationship between SO2 and other pollutant emissions from
uidized-bed combustion. Symp (Int) Combust 1998;27:3093101.
[14] Miccio F, Löer G, Wargadalam VJ, Winter F. The inuence of SO2 level and
operating conditions on NOx and N2O emissions during uidised bed combustion of
coals. Fuel 2001;80:155566.
[15] Alzueta MU, Bilbao R, Glarborg P. Inhibition and sensitization of fuel oxidation by
SO2. Combust Flame 2001;127:223451.
[16] Weng W, Chen S, Wu H, Glarborg P, Li Z. Optical investigation of gas-phase KCl/
KOH sulfation in post ame conditions. Fuel 2018;224:4618.
[17] Roesler JF, Yetter RA, Dryer FL. The inhibition of the CO/H20/02 reaction by trace
quantities of HCl. Combust Sci Technol 1992;82:87100.
[18] Wu D, Wang Y, Wei X, Li S, Guo X. Experimental investigation and numerical si-
mulation of CO oxidation with HCl addition. Fuel 2016;179:2218.
[19] Julien S, Brereton CMH, Lim CJ, Grace JR, Anthony EJ. The eect of halides on
emissions from circulating uidized bed combustion of fossil fuels. Fuel
1996;75:165563.
[20] Wang J, Anthony EJ. CO oxidation and the inhibition eects of halogen species in
uidised bed combustion. Combust Theor Model 2009;13:10519.
[21] Weng WB, Borggren J, Li B, Alden M, Li ZS. A novel multi-jet burner for hot ue
gases of wide range of temperatures and compositions for optical diagnostics of
solid fuels gasication/combustion. Rev Sci Instrum 2017;88.
[22] Borggren J, Weng WB, Hosseinnia A, Bengtsson PE, Alden M, Li ZS. Diode laser-
based thermometry using two-line atomic uorescence of indium and gallium. Appl
Phy B-Lasers Opt 2017;123.
[23] Gao Q, Weng WB, Li B, Li ZS. Quantitative and spatially resolved measurement of
atomic potassium in combustion using diode laser. Chin Phys Lett 2018;35.
[24] Weng W, Brackmann C, Leer T, Aldén M, Li Z. Ultraviolet absorption cross sec-
tions of KOH and KCl for nonintrusive species-specic quantitative detection in hot
ue gases. Anal Chem 2019;91:471926.
[25] Qu ZC, Steinvall E, Ghorbani R, Schmidt FM. Tunable diode laser atomic absorption
spectroscopy for detection of potassium under optically thick conditions. Anal
Chem 2016;88:375460.
[26] Weng WB, Leer T, Brackmann C, Alden M, Li ZS. Spectrally resolved ultraviolet
(UV) absorption cross-sections of alkali hydroxides and chlorides measured in hot
ue gases. Appl Spectrosc 2018;72:138895.
[27] Luque J, Crosley DR. LIFBASE: database and spectral simulation program (Version
1.5). SRI International Report MP 99-009; 1999.
[28] Luque J, Crosley DR. LIFBASE: database and spectral simulation program (Version
1.5); 1999.
[29] Design R. CHEMKIN-PRO 15131 San Diego; 2013.
[30] Weng W, Zhang Y, Wu H, Glarborg P, Li Z. Optical measurement of KOH, KCl and K
for quantitative K-Cl chemistry in thermochemical conversion processes, Submitted
to Fuel.
[31] Smith GP, Golden DM, Frenklach M, Moriarty NW, Eiteneer B, Goldenberg M,
Bowman CT, Hanson RK, Song S, Gardiner WC, VVL Jr, Qin Z. GRI-Mech 3.0.
http://www.me.berkeley.edu/gri_mech/.
[32] Glarborg P, Marshall P. Mechanism and modeling of the formation of gaseous alkali
sulfates. Combust Flame 2005;141:2239.
[33] Hindiyarti L, Frandsen F, Livbjerg H, Glarborg P, Marshall P. An exploratory study
of alkali sulfate aerosol formation during biomass combustion. Fuel
2008;87:1591600.
[34] Li B, Sun Z, Li Z, Aldén M, Jakobsen JG, Hansen S, et al. Post-ame gas-phase
sulfation of potassium chloride. Combust Flame 2013;160:95969.
[35] Song Y, Hashemi H, Christensen JM, Zou C, Haynes BS, Marshall P, et al. An ex-
ploratory ow reactor study of H2S oxidation at 30100 bar. Int J Chem Kinet
2017;49:3752.
[36] Pelucchi M, Frassoldati A, Faravelli T, Ruscic B, Glarborg P. High-temperature
chemistry of HCl and Cl2. Combust Flame 2015;162:2693704.
[37] Sorvajärvi T, Viljanen J, Toivonen J, Marshall P, Glarborg P. Rate constant and
thermochemistry for K + O2 + N2 = KO2 + N2. J Phys Chem A
2015;119:332936.
[38] Kaskan WE. The reaction of alkali atoms in lean ames. Symp (Int) Combust
1965;10:416.
[39] McEwan M, Phillips L. Dissociation energy of NaO2. Trans Faraday Soc 1966;62.
[40] Carabetta R, Kaskan WE. Oxidation of sodium, potassium, and cesium in ames. J
Phys Chem 1968;72:24839.
[41] Hynes AJ, Steinberg M, Schoeld K. The chemical kinetics and thermodynamics of
sodium species in oxygen-rich hydrogen ames. J Chem Phys 1984;80:258597.
[42] Mortensen MR, Hashemi H, Wu H, Glarborg P. Modeling post-ame sulfation of KCl
and KOH in bio-dust combustion with full and simplied mechanisms. Fuel
2019;258:116147.
[43] Perry RA, Miller JA. An exploratory investigation of the use of alkali metals in
nitrous oxide control. Int J Chem Kinet 1996;28:21734.
T. Berdugo Vilches, et al. Fuel 273 (2020) 117762
8
... The modeling predictions for the sulfation of KCl by H 2 SO 4 in gaseous interactions were performed with the plug flow reactor configuration in the Chemkin PRO. The reaction mechanism and thermodynamic data, available as Supporting Information, were largely drawn from the recent work of Chanpirak et al. 61 It was based on the work of Glarborg and Marshal, 18 Hindiyarti et al., 20 Weng et al., 58 and Berdugo Vilchez et al., 62 consisting of subsets for potassium, 21,58 chlorine, 63,64 and sulfur 65,66 chemistry as well as the condensation reaction of K 2 SO 4 (g). 21 Modeling predictions compared favorably with experiments in laboratory flow reactors and entrained-flow reactors 16,19,22 as well as data from other units. ...
... The experimental results were interpreted in terms of the detailed chemical kinetic models of Berdugo Vilchez et al. 62 and Chanpirak et al. 61 In both mechanisms, the reaction SO 3 + 2H 2 O ⇌ H 2 SO 4 + H 2 O was added to describe decomposition of sulfuric acid. Modeling predictions were conducted using the temperature profile of the isothermal and cooling zones. ...
... The modeling simulations shown in Figure 4 were conducted with the mechanisms of Berdugo Vilchez et al. 62 and Chanpirak et al., 61 respectively, taking into account the isothermal zone and the cooling region. Because part of the KCl was in condensed form and thus unavailable for reaction below 1000 K, modeling predictions are only shown for experiments with set temperatures above 1000 K, and the preheating section was not considered in the calculations. ...
Article
The sulfation of gaseous KCl by H2SO4 in the presence of O2 and H2O was investigated in a laboratory laminar flow reactor at atmospheric pressure and temperatures ranging from 873 to 1523 K. The degree of sulfation was characterized by analysis of the elemental composition (Cl, S, and K) of the submicrometer particles captured on a filter downstream of the reactor. Furthermore, concentrations of HCl and SO2 in the outlet were measured. The experimental results were interpreted in terms of a detailed chemical kinetic model for the K/S/Cl chemistry. The degree of KCl sulfation depends on the reaction temperature and the reactant concentrations. The results show that H2SO4 is effective in sulfating KCl to K2SO4 at temperatures in the range 1073–1273 K, in particular at high H2SO4/KCl ratios. At temperatures of 1373 K and above, the sulfation efficiency decreases rapidly because alkali sulfate formation is no longer thermodynamically favored. The kinetic analysis shows that reaction is initiated by decomposition of sulfuric acid: H2SO4 + H2O ⇄ SO3 + H2O + H2O. Then KCl reacts directly with SO3 to form the intermediate KSO3Cl, which is further converted via KHSO4 to gaseous K2SO4. As the gas is cooled downstream of the isothermal zone, potassium sulfate nucleates homogeneously to form an aerosol, promoting further sulfation in the gas phase. The simulations are in satisfactory agreement with the experimental results, predicting correctly the temperature window for the process. However, the model overpredicts the SO2 yield at high temperature, indicating that the K/S interactions under these conditions are not fully understood. The results are important, both for validation of models of K/Cl/S interactions in combustion and for characterization of the sulfation of KCl by sulfuric acid to evaluate the potential of the sulfur recirculation process and identify the optimum process conditions.
... UV-vis spectrum of KOH sample contains: (i) two well-defined absorption bands at wavelength values of 239 nm and 287 nm and (ii) peaks of low intensity located at 322 nm and 358 nm. The bands within 230-290 nm were assigned to electronic transitions OH − → K + between K cation and hydroxyl anion in KOH phase, while those between 320 and 360 nm denote electron density distribution due to CO 2− 3 → 2K + s charge transfers in K 2 CO 3 [36][37][38]. These findings agree well with the XRD method, where bare KOH was established as partially carbonized. ...
Article
Potassium hydroxide (KOH) immobilized on rapana shells-based anorthite (ART) and waste coal fly ash (CFA) was investigated as a catalyst (KOH/ART and KOH/CFA) for sunflower oil methanolysis (biodiesel production) for the first time. A number of physicochemical methods such as SBET, XRD, FT-IR, UV–vis, XPS, SEM, TEM and TPD-CO2 have been used for characterization of KOH/ART and KOH/CFA catalysts. Important kinetic parameters (rate constant, activation energy, pre-exponential factor) and thermodynamic functions (enthalpy, entropy and Gibbs free energy) for the reaction of biodiesel production were also evaluated. In addition, a plausible reaction mechanism has been offered. Textural and surface measurements revealed the presence of dispersed alkaline particles (KNaCO3, K2CO3, [Ca,K,Na][Al,Si]4O8, K4.04Ca0.66Na0.34Al5.7Si10.3O32) on KOH/ART and KOH/CFA surface due to strong active phase-support interaction. As a result, higher basicity and catalytic performance for KOH/ART with respect to KOH/CFA in the process of sunflower oil methanolysis has been established.
Article
Full-text available
Biomass conversion with carbon capture and storage (Bio‐Energy CCS; BECCS) is one of the options considered for mitigating climate change. In this paper, the carbon capture technology chemical‐looping combustion (CLC) is examined in which the CO2 is produced in a stream separate from the combustion air. A central research topic for CLC is oxygen carriers; solid metal oxides that provide oxygen for the conversion process. Biomass and waste‐derived fuels contain reactive ash compounds, such as potassium, and interactions between the oxygen carrier and the ash species are critical for the lifetime and performance of the oxygen carrier. This work develops and demonstrates an improved method for studying the interactions between ash species and oxygen carriers. The method uses a lab‐scale reactor operating under fluidized conditions, simulating CLC batch‐wise by switching between feed gas. The novelty of the setup is the integrated system for feeding solid particles of ash model compounds, enabling the simulation of ash species accumulating in the bed. Ilmenite is a benchmark oxygen carrier for solid fuel conversion and was used in this study to evaluate the method using K2CO3 as a model ash compound. Experiments were done at 850 and 950°C. Methane conversion in CLC cycles and fluidization was evaluated with gas analysis and pressure drop measurements. Scanning electron microscopy and energy dispersive X‐ray spectroscopy (SEM‐EDS) and X‐ray diffraction (XRD) analysis of bed particles were done after the experiments to establish changes in the morphology and composition of the ilmenite. The method for feeding the ash model compound was concluded to be satisfactory. At 950°C, K accumulated in the particles forming K‐titanates and agglomeration was enhanced with K2CO3 addition. The agglomeration mechanism was solid‐state sintering between the Fe‐oxides forming on the particle surfaces. The bed defluidized at 950°C, but no such effect was seen at 850°C. The method is suitable for studying the Fe‐Ti‐K system with ilmenite and potassium without the influence of other ash species. © 2023 The Authors. Greenhouse Gases: Science and Technology published by Society of Chemical Industry and John Wiley & Sons Ltd.
Article
Full-text available
This paper presents the main experiences gained and conclusions drawn from the demonstration of a first-of-its-kind wood-based biomethane production plant (20-MW capacity, 150 dry tonnes of biomass/day) and 10 years of operation of the 2–4-MW (10–20 dry tonnes of biomass/day) research gasifier at Chalmers University of Technology in Sweden. Based on the experience gained, an elaborated outline for commercialization of the technology for a wide spectrum of applications and end products is defined. The main findings are related to the use of biomass ash constituents as a catalyst for the process and the application of coated heat exchangers, such that regular fluidized bed boilers can be retrofitted to become biomass gasifiers. Among the recirculation of the ash streams within the process, presence of the alkali salt in the system is identified as highly important for control of the tar species. Combined with new insights on fuel feeding and reactor design, these two major findings form the basis for a comprehensive process layout that can support a gradual transformation of existing boilers in district heating networks and in pulp, paper and saw mills, and it facilitates the exploitation of existing oil refineries and petrochemical plants for large-scale production of renewable fuels, chemicals, and materials from biomass and wastes. The potential for electrification of those process layouts are also discussed. The commercialization route represents an example of how biomass conversion develops and integrates with existing industrial and energy infrastructures to form highly effective systems that deliver a wide range of end products. Illustrating the potential, the existing fluidized bed boilers in Sweden alone represent a jet fuel production capacity that corresponds to 10% of current global consumption.
Article
Full-text available
compact optical setup for quantitative and spatially resolved measurement of atomic alkali concentration in combustion is demonstrated. Tunable diode laser absorption spectroscopy and laser-induced fluorescence are combined using a single continuous wave diode laser to measure the line-integration concentration and the relative distribution simultaneously, thereby obtaining the absolute concentration distribution along the laser beam. The results indicate the good performance of this method for one-dimensional quantitative measurement.
Article
Full-text available
A robust and relatively compact calibration-free thermometric technique using diode lasers two-line atomic fluorescence (TLAF) for reactive flows at atmospheric pressures is investigated. TLAF temperature measurements were conducted using indium and, for the first time, gallium atoms as temperature markers. The temperature was measured in a multi-jet burner running methane/air flames providing variable temperatures ranging from 1600 to 2000 K. Indium and gallium were found to provide a similar accuracy of ~ 2.7% and precision of ~ 1% over the measured temperature range. The reliability of the TLAF thermometry was further tested by performing simultaneous rotational CARS measurements in the same experiments.
Article
Potassium and chlorine chemistry at high temperature is of great importance in biomass utilization through thermal conversion. In well-defined hot environments, we performed quantitative measurements of main potassium species, i.e., potassium hydroxide (KOH), potassium chloride (KCl) and K atoms, and the important radical OH. The concentrations of KOH, KCl and OH radicals were measured through a newly developed UV absorption spectroscopy technique. Quantitative measurements of potassium atoms were performed using tunable diode laser absorption spectroscopy at the wavelength of 404.4 and 769.9 nm to cover a wide concentration dynamic range. The reaction environment was provided by a laminar flame burner, covering a temperature range of 1120–1950 K and global fuel-oxygen equivalence ratios from 0.67 to 1.32. Potassium and chlorine were introduced into the combustion atmosphere by atomized K2CO3 or KCl water solution fog. The experimental results were compared to modeling predictions to evaluate a detailed K-Cl mechanism. For most cases, the experimental and simulation results were in reasonable agreement. However, the over-prediction of K atom concentration at low temperature fuel-rich condition and the overall under-prediction of KCl concentration call for further investigation. It was demonstrated that the optical methods and the well-defined hot environments could provide quantitative investigations widely applicable to different homogeneous reactions in thermochemical conversion processes, and in evaluation of corresponding reaction mechanisms with reliable data.
Article
The gas-phase interaction between alkali volatiles and sulphur oxides has important implications for deposition and corrosion in combustion of biomass. In the present study, gas-phase transformation of KOH and KCl in the post-flame zone in bio-dust combustion has been studied by detailed chemical kinetic modeling for a woody and a herbaceous biomass, respectively. For both biomasses K > Cl on a molar basis, and KOH is the major alkali species at high temperature. The modeling indicates that KOH is readily converted to K2SO4 in the presence of SO2 at temperatures below 1500 K. Below 1100 K, gaseous K2SO4 nucleates homogeneously, promoting further sulfation. For the woody biomass, where K < Cl + 2S, the sulfation process is kinetically limited due to the competition with chlorination. For the herbaceous biomass, with K > Cl + 2S, the excess KOH facilitates internal equilibration among the alkali species. For both biomasses, the KCl concentration remains constant during sulfation, because any KCl consumed is rapidly replenished by the reaction KOH + HCl ⇄ KCl + H2O. The consequence is that the sulfation process under the investigated conditions does not help to remove the main corrosive agent, KCl. For use in CFD, a skeletal model for the gas-phase sulfation of KCl and KOH has been developed, based on systematic reduction of the detailed chemical kinetic model. However, for fuels with excess K compared to Cl + 2S, a simple equilibrium calculation may yield satisfactory results.
Article
Understanding of potassium chemistry in energy conversion processes supports the development of complex biomass utilization with high efficiency and low pollutant emissions. Potassium exists mainly as potassium hydroxide (KOH), potassium chloride (KCl) and atomic potassium (K) in combustion and related thermochemical processes. We report, for the first time, the measurement of the UV absorption cross-sections of KOH and KCl at temperatures between 1300 and 1800 K using a newly developed method. Using the spectrally-resolved UV absorption cross-sections, the concentrations of KOH and KCl were measured simultaneously. Additionally, we measured the concentrations of K atom using tunable diode laser absorption spectroscopy both at 404.4 nm and 769.9 nm. The 404.4 nm line was utilized to expand the measurement dynamic range to higher concentrations. A constant amount of KCl was seeded into premixed CH4/air flames with equivalence ratios varied from 0.67 to 1.32, and the concentrations of KOH, KCl and K atoms in the hot flue gas were measured nonintrusively. The results indicate that these techniques can provide comprehensive data for quantitative understanding of the potassium chemistry in biomass combustion/gasification.
Article
A counter-flow reactor setup was designed to investigate the gas-phase sulfation and homogeneous nucleation of potassium salts. Gaseous KOH and KCl were introduced into the post-flame zone of a laminar flat flame. The hot flame products mixed in the counter-flow with cold N2, with or without addition of SO2. The aerosols formed in the flow were detected through Mie scattering of a 355 nm laser beam. The temperature distribution of the flow was measured by molecular Rayleigh scattering thermometry. From the temperature where nucleation occurred, it was possible to identify the aerosols formed. Depending on the potassium speciation in the inlet and the presence of SO2, they consisted of K2SO4, KCl, or K2CO3, respectively. The experiments showed that KOH was sulphated more readily than KCl, resulting in larger quantities of aerosols. The sulfation process in the counter-flow setup was simulated using a chemical kinetic model including a detailed subset for the Cl/S/K chemistry. Similar to the experimental results, much more potassium sulfate was predicted when seeding KOH compared to seeding KCl. For both KOH and KCl, sulfation was predicted to occur primarily through the reactions among atomic K, O2 and SO2, forming KHSO4 and K2SO4. The higher propensity for sulfation of KOH compared to KCl was mostly attributed to the lower thermal stability of KOH, facilitating formation of atomic K. According to the model, sulfation also happened through SO3, especially for KCl (KCl → KSO3Cl → K2SO4).
Article
Spectrally resolved ultraviolet (UV) absorption cross-sections of gas-phase sodium chloride (NaCl), potassium hydroxide (KOH), and sodium hydroxide (NaOH) were measured, for the first time, in hot flue gases at different temperatures. Homogenous gas-phase NaCl, KCl (potassium chloride), NaOH, and KOH at temperatures 1200 K, 1400 K, 1600 K, and 1850 K were prepared in the post-flame zone of laminar flames by seeding nebulized droplets out of aqueous solution of corresponding alkali species. The amount of droplets seeded into the flame was kept constant, so the relative concentration of different alkali species can be derived. The broadband UV absorption cross-section of KCl vapor reported by Leffler et al. was adopted to derive the absorption cross-section curves of NaCl, NaOH, and KOH with the corresponding measured spectrally resolved absorbance spectra. No significant changes in the spectral structures in the absorption cross-sections were found as the temperature varied between 1200 K and 1850 K, except for NaOH at around 320 nm. The difference between the absorption spectral curves of alkali chlorides and hydroxides is significant at wavelengths above 300 nm, which thus can be used to distinguish and obtain the concentrations of alkali chlorides and hydroxides in the broadband UV absorption measurements.
Article
A novel multi-jet burner was built to provide one-dimensional laminar flat flames with a wide range of variable parameters for multipurpose quantitative optical measurements. The burner is characterized by two independent plenum chambers, one supporting a matrix of 181 laminar jet flames and the other supporting a co-flow from a perforated plate with small holes evenly distributed among the jets. A uniform rectangular burned gas region of 70 mm × 40 mm can be generated, with a wide range of temperatures and equivalence ratios by controlling independently the gas supplies to the two plenum chambers. The temperature of the hot gas can be adjusted from 1000 K to 2000 K with different flame conditions. The burner is designed to seed additives in gas or liquid phase to study homogeneous reactions. The large uniform region can be used to burn solid fuels and study heterogeneous reactions. The temperature was measured using two-line atomic fluorescence thermometry and the temperature profile at a given height above the burner was found to be flat. Different types of optical diagnostic techniques, such as line of sight absorption or laser-induced fluorescence, can be easily applied in the burner, and as examples, two typical measurements concerning biomass combustion are demonstrated.
Article
Hydrogen sulfide oxidation experiments were conducted in O2/N2 at high pressure (30 and 100 bar) under oxidizing and stoichiometric conditions. Temperatures ranged from 450 to 925 K, with residence times of 3–20 s. Under stoichiometric conditions, the oxidation of H2S was initiated at 600 K and almost completed at 900 K. Under oxidizing conditions, the onset temperature for reaction was 500–550 K, depending on pressure and residence time, with full oxidization to SO2 at 550–600 K. Similar results were obtained in quartz and alumina tubes, indicating little influence of surface chemistry. The data were interpreted in terms of a detailed chemical kinetic model. The rate constants for selected reactions, including SH + O2 ⇄ SO2 + H, were determined from ab initio calculations. Modeling predictions generally overpredicted the temperature for onset of reaction. Calculations were sensitive to reactions of the comparatively unreactive SH radical. Under stoichiometric conditions, the oxidation rate was mostly controlled by the SH + SH branching ratio to form H2S + S (promoting reaction) and HSSH (terminating). Further work is desirable on the SH + SH recombination and on subsequent reactions in the S2 subset of the mechanism. Under oxidizing conditions, a high O2 concentration (augmented by the high pressure) causes the termolecular reaction SH + O2 + O2 HSO + O3 to become the major consumption step for SH, according to the model. Consequently, calculations become very sensitive to the rate constant and product channels for the H2S + O3 reaction, which are currently not well established.