You are on page 1of 15

ARTICLE IN PRESS CNF:6965

JID:CNF AID:6965 /FLA [mNY4a; v 1.89; Prn:5/03/2008; 14:39] P.1 (1-15)

Combustion and Flame ••• (••••) •••–•••


www.elsevier.com/locate/combustflame

Scaling of NOx emissions from a laboratory-scale mild


combustion furnace
G.G. Szegö ∗ , B.B. Dally, G.J. Nathan
School of Mechanical Engineering, The University of Adelaide, South Australia 5005, Australia
Received 10 September 2007; received in revised form 2 February 2008; accepted 7 February 2008

Abstract
A systematic experimental campaign has been carried out to investigate the scaling of NOx emissions from a
moderate or intense low-oxygen dilution (MILD) combustion furnace operating with a parallel jet burner system
in which the reactants and the exhaust ports are all mounted on the same wall. Its maximum capacity was 20 kW
from the fuel and 3.3 kW from air preheat, with a turndown ratio of 1:3. The burner system was configured to
achieve high dilution of the incoming reactants. A comprehensive data set comprising 191 global measurements of
temperature and exhaust gas emissions is presented, together with temperature contours on the furnace centerline
plane. It was found that, although heat extraction, air preheat, excess air, firing rate, dilution, and fuel type all
affect global NOx emissions, they do not control NOx scaling. The combined effects of these global parameters
can be ultimately characterized by a furnace temperature and a global residence time. A temperature–time scaling
approach, previously reported for open jet diffusion flames, proved to be a useful tool for comparison of NOx
emissions from highly diluted furnace environments regardless of the furnace/burner geometries. Regression-
based predictions found the characteristic temperature to correlate with 85% of the data with an accuracy of only
±50%. The leading-order approach also showed that the jet exit Froude number is of limited value for NOx scaling
in the MILD regime. Because of the weak dependence on temperature observed in the data and the moderate
magnitude of the measured temperatures, it is deduced that the prompt-NO and/or N2 O-intermediate pathways
are of significance comparable to that of the thermal-NO pathway. The analysis also suggests that NOx formation
is controlled neither by kinetics nor by mixing, and hence the conditions inside this furnace approach or span the
range in which Damköhler numbers are of order unity, Da = O(1).
© 2008 The Combustion Institute. Published by Elsevier Inc. All rights reserved.

Keywords: MILD furnace; Moderate temperatures; NOx scaling

1. Introduction change, have intensified discussions about the com-


bustion of fossil and alternative fuels. Combustion-
Rising concentrations of atmospheric greenhouse generated CO2 emissions account for about 80% of
gases, which are widely accepted to lead to climate anthropogenic greenhouse gases [1], so their mitiga-
tion is vital, and an aim of policies around the world.
* Corresponding author. Fax: +61 8 8303 4367. Improving energy efficiency is often the most cost-
E-mail address: george.szego@adelaide.edu.au effective way to reduce CO2 emissions [2]. A wide
(G.G. Szegö). range of low- or zero-carbon-emission technologies
0010-2180/$ – see front matter © 2008 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
doi:10.1016/j.combustflame.2008.02.001

Please cite this article in press as: G.G. Szegö et al., Scaling of NOx emissions from a laboratory-scale mild combustion
furnace, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.02.001
ARTICLE IN PRESS CNF:6965
JID:CNF AID:6965 /FLA [mNY4a; v 1.89; Prn:5/03/2008; 14:39] P.2 (1-15)
2 G.G. Szegö et al. / Combustion and Flame ••• (••••) •••–•••

are already available [3]. However, the abatement


of pollutants often comes at the price of efficiency
losses. Moderate or intense low-oxygen dilution
(MILD) combustion [4] is a credible candidate to si-
multaneously meet thermal efficiency needs and pol-
lutant emission restrictions. Under MILD conditions,
reactants are highly diluted with hot combustion prod-
ucts, causing reactions to occur in a distributed reac-
tion zone with reduced peak temperatures [5]. As a
consequence, the temperature distribution is nearly
uniform, the net radiation flux can be enhanced, and
pollutant emissions, NOx in particular, are lower than
those from conventional flames.
Over the past few decades, MILD combustion
technology has been implemented at full scale in var-
ious industrial sectors [6] and tested at pilot scale in
other applications [7–10] and for other fuel types [11].
Nevertheless, despite considerable industrial success,
important aspects of MILD combustion remain un-
known. For example, little is known about the scaling
relations of NOx emissions in complex recirculating
flows with low oxygen concentrations at moderate
temperature ranges (800 ◦ C < T < 1400 ◦ C), typi-
cally found in MILD conditions.
The scaling of NOx has been comprehensively in-
vestigated in unconfined nonpremixed turbulent jet Fig. 1. Schematic diagram of MILD combustion furnace and
flames [12], and also given considerable attention parallel jet burner system.
in enclosed furnace environments with conventional
combustion systems [13–16]. These studies have examine the NOx emissions from a laboratory-scale
found a strong correlation of normalized NOx emis- MILD combustion furnace (MCF), operating with a
sion indices with fuel jet exit Froude number and of parallel jet burner system in which the reactants and
NOx formation rates with global temperatures and the exhaust ports are all mounted on the same wall.
residence times. These relationships are useful, not The maximum furnace capacity is about 20 kW from
only for understanding which geometrical or opera- the fuel and 3.3 kW from air preheat. First, results of
tional parameters control pollutant formation, but also thermal field measurements are presented to provide
for assessing the relative importance of the alternative some insight into the major NOx formation and/or de-
NOx formation mechanisms. struction regions. Then, the effects of fuel dilution on
More specifically related to the MILD regime, So- global NOx emissions are discussed. Finally, a com-
biesiak et al. [17] obtained an empirical correlation prehensive experimental data set, which covers a va-
for NOx emissions as a function of excess air, air riety of operating conditions, is used to investigate the
preheat temperature, burner turndown ratio, and ex- scaling of NOx emissions with global parameters.
haust gas temperature. Although relatively few exper-
imental data points were available for the analysis, a
power-law type expression was used to show that the 2. Experimental arrangement and methodology
exhaust temperature term had the strongest effect on
NOx emissions. Other researchers [18,19] attempted 2.1. MILD combustion furnace and burner
to relate NOx emission trends to burner parameters,
such as fuel injection angle and fuel-to-air velocity ra- Fig. 1 shows a schematic diagram of the MILD
tios. The aforementioned MILD combustion studies combustion furnace and the parallel jet burner config-
focused on minimizing NOx emissions of a particu- uration built for this project. The combustion cham-
lar burner design and led to the patenting [20–22] and ber has a square cross section of 280 × 280 mm2
commercialization of the technology. and a height of 585 mm. It is well insulated, with
To the best of our knowledge, the results of NOx four layers of 38-mm-thick high-temperature ceramic
scaling relations have not been previously reported fiberboards that allow only about 20% of heat to be
for MILD combustion conditions, either in enclosed lost through the walls. This assists in the establish-
or in open flame systems. The current work aims to ment and stability of the MILD regime and results in

Please cite this article in press as: G.G. Szegö et al., Scaling of NOx emissions from a laboratory-scale mild combustion
furnace, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.02.001
ARTICLE IN PRESS CNF:6965
JID:CNF AID:6965 /FLA [mNY4a; v 1.89; Prn:5/03/2008; 14:39] P.3 (1-15)
G.G. Szegö et al. / Combustion and Flame ••• (••••) •••–••• 3

a warm-up time of about 3 h from a cold state un- verted to temperature units with a USB-TC data log-
til steady-state operation. The inner surfaces of the ger. According to the manufacturer’s specifications,
combustion chamber were coated with a commercial the uncertainties associated with linearization, cold-
product, “Kaowool J-Coat” rigidizer, to improve the junction compensation, and system noise produce a
resistance to erosion caused by high gas velocities. maximum error of ±1.07 ◦ C in a measurement range
The MCF has full optical access through five open- from 250 to 1768 ◦ C. This represents only 0.13% er-
ings (120 × 76 mm2 ) equally spaced along the ver- ror in the worst case.
tical direction on three sides of the furnace. These Eighteen thermocouple ports (see Fig. 1) were
openings can accommodate interchangeable insulat- equally spaced in the vertical direction, through
ing window plugs or UV-grade fused silica windows. which a thermocouple could be traversed. Typically,
The chamber is slightly pressurized to prevent air temperatures were monitored simultaneously by fixed
ingress, and high-temperature gaskets were used to thermocouples in two regions of interest, one at the
seal all surface joints for safety. A U-shaped cool- top section and one at the bottom section of the com-
ing loop with variable heat exchange area was used bustion chamber. The center of the air nozzle exit
to control the heat load. The heat exchanger can be plane is defined to be the origin of a Cartesian co-
inserted through any of the five window openings (po- ordinate system (x, y, z). The top temperature was
sitions A1–A5). measured on the furnace centerline (0, 0, 542.5) and
The configuration of reactants and exhaust ports was denoted as the furnace reference temperature.
was optimized using a computational fluid dynam- The bottom temperature was measured flush with
ics (CFD) modeling study [23]. The furnace was de- the inner surface (0, 140, 42.5) and was denoted as
signed for a maximum capacity of 20 kW from the the reference wall temperature. Temperature profiles
fuel and 3.3 kW from air preheat with a turndown ra- were also measured for a selected baseline case. The
tio of 1:3. The air was preheated with a Leister LE thermocouple probes were traversed in the radial di-
3000 electrical air heater that has a built-in poten- rection (y-direction) across different planes at each
tiometer for air temperature control. The burner con- hole to provide a total of 72 points.
sists of a single air nozzle on the axis of the furnace, The correction to the temperature measurements
and four exhaust ports and four fuel ports arranged for radiation was determined from a steady-state en-
symmetrically in ring patterns (see Fig. 1). The dis- ergy balance on the thermocouple bead surface. The
tance between the center of the air nozzle and the actual gas temperature (Tg ) was determined from the
centers of the exhaust ports and fuel nozzles are 55 measured temperature (Ttcb ) and net radiation flux
and 110 mm, respectively. In this arrangement the in- between diffuse-gray surfaces in an enclosure accord-
coming air stream is separated from the fuel streams ing to
by outgoing exhaust gases, much as in the fuel direct 
injection (FDI) concept introduced by Nakamachi et Atcb εtcb N 4 4
w=1 σ G1w (Ttcb − Tw )
Tg = Ttcb + , (1)
al. [24] and in the IFRF burner design [25]. This h
ensures that the reactants are diluted with products where Atcb is the thermocouple bead surface area,
before any combustion reaction can occur. The final εtcb is the thermocouple bead surface emissivity, N
straight sections of the air and fuel supply lines have is the total number of surfaces (i.e., walls) in the en-
length-to-diameter ratios of L/D ≈ 10 and L/d ≈ 18 closure, σ is the Stefan–Boltzmann constant, G1w is
(id = 2 mm), respectively. the fraction of radiation emitted by the thermocouple
bead that is incident on a particular wall surface w
2.2. Temperature and global emission measurements in the furnace enclosure and is absorbed [26], Tw is
the surface temperature of wall w in the enclosure,
Time-averaged temperatures were measured with and h is the convection coefficient calculated for a
bare, fine-wire type R (Pt/Pt–13% Rh) thermocou- sphere [27]. Catalytic and oxidation effects were ne-
ples of 254-µm-diameter wires and 1.2-mm-diameter glected, as well as conductive heat transfer along the
beads under steady-state conditions. This is sufficient ceramic sheath. The G1w correction factor was cal-
for reliable measurements of the mean temperature culated for each probe position. The method led to a
after correction for radiation, given the low temper- maximum correction of 9% and a typical correction
ature fluctuations in the MILD regime, and provides of less than 2% between raw and actual gas tempera-
satisfactory spatial resolution while being sufficiently tures, so that the total uncertainty is within ±1.2%.
robust to avoid probe breakage. For accurate posi- Exhaust gas composition was measured inside
tioning, the thermocouple wires were supported by the exhaust pipe with a TESTO 350XL portable
a 6-mm-diameter rigid ceramic sheath, with the final gas analyzer. The gas analyzer was equipped with
5 mm left exposed. The differential voltage signal was electrochemical sensors with nominal accuracies,
continuously sampled at 1 Hz and automatically con- respectively, for CO (±5%: 0–2000 ppm, ±10%:

Please cite this article in press as: G.G. Szegö et al., Scaling of NOx emissions from a laboratory-scale mild combustion
furnace, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.02.001
ARTICLE IN PRESS CNF:6965
JID:CNF AID:6965 /FLA [mNY4a; v 1.89; Prn:5/03/2008; 14:39] P.4 (1-15)
4 G.G. Szegö et al. / Combustion and Flame ••• (••••) •••–•••

2001–10000 ppm), NO (±5 ppm: 0–99 ppm), NO2


(±5 ppm: 0–99.9 ppm), O2 (±0.8%: 0–25%), and
UHC (<400 ppm: 100–4000 ppm CH4 ) and with
a nondispersive infrared (NDIR) sensor for CO2
(±0.3% m.v. + 0.44: 0–25%). The analyzer was
checked with a calibration gas to yield total NOx
emission accuracies better than ±4 ppm for the range
from 0 to 70 ppm, in which all present data lie.
The combustion products were cooled and filtered
to remove moisture and particles before reaching the
sensing cells. Total NOx emission (NO + NO2 ) con-
centrations are reported by volume on a dry basis
corrected to 3% O2 concentration. Emission indices
are expressed as the ratio of pollutant formed to the
mass of fuel input [28]. The emission indices were
determined from measured species concentrations as
XNO MWNO + XNO2 MWNO2
EINOx [g/kg fuel] =
(XCO + XCO2 )MWCx Hy
× 1000, (2)
XCO MWCO
EICO [g/kg fuel] =
(XCO + XCO2 )MWCx Hy
× 1000, (3)
where EINOx is the NOx emission index, Xi is the
mole fraction of species i, and MW is the molec-
ular weight of species i or of the hydrocarbon fuel
(Cx Hy ). For most operating conditions, unburned hy-
drocarbons (UHC) were either below the detection
limit of 100 ppm or below the minimum O2 concen-
tration (≈2%) required for a reliable measurement.
Therefore, for consistency, UHC was neglected as a
possible source of carbon in the formulation. Due
to the short length of the flue gas sampling probe
and hose (≈2.5 m), the manufacturer reports that the
amount of NO2 absorbed in the condensate trap is in-
significant. Consequently, both measured components
of NOx can be considered to be reliable.

2.3. Operating conditions

The parametric study investigated a total of 191 Fig. 2. Schematic diagram of heat exchanger insertion posi-
measurement points under various test conditions. In tions.
all cases, thermal equilibrium was ensured and sta-
ble MILD combustion was achieved, based on the
requirement of no visible flame and NOx emission from 0.67 to 0.98, which is equivalent to an excess air
levels being below 70 ppm throughout the experi- range of 33–2%. The location of the heat exchanger
ment. The oxidant was air, whose temperature was was varied from the top window A1 (centerline at
varied from ambient to 780 ◦ C. Either commercial z = 524.5 mm) to the bottom window A4 (centerline
natural gas (NG) or liquefied petroleum gas (LPG) at z = 176.5 mm) and its exposed surface area varied
was used as the fuel to provide thermal inputs over from 0.015 m2 for 25% insertion to 0.06 m2 for 100%
the range 7.5–20 kW. Table 1 shows typical compo- insertion (Fig. 2). The dilution of the fuel stream was
sition and properties of both fuels. Two alternative- varied, using either N2 or CO2 , over the range 0–76%
sized fuel nozzles were employed, 2.0 or 3.0 mm by mass. Table 2 presents a summary of all test con-
in diameter. The global equivalence ratio (φ) ranged ditions.

Please cite this article in press as: G.G. Szegö et al., Scaling of NOx emissions from a laboratory-scale mild combustion
furnace, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.02.001
ARTICLE IN PRESS CNF:6965
JID:CNF AID:6965 /FLA [mNY4a; v 1.89; Prn:5/03/2008; 14:39] P.5 (1-15)
G.G. Szegö et al. / Combustion and Flame ••• (••••) •••–••• 5

Table 1
Typical fuel properties and composition
Fuel LHVb (A/F)st c MWd N2 e CO2 e CH4 e C2 H6 e C3 H8 e C4 H10 e C5 H12 e C6 H14 e
NGa 51.154 17.11 17.71 1.278 2.084 91.36 4.364 0.62 0.20 0.055 0.04
LPGa 45.993 15.66 44.07 0.73 0.05 – 1.12 96.32 1.78 – –
a Gas analysis provided by Origin Energy Australia.
b Lower heating value (MJ/kg).
c Stoichiometric air-to-fuel ratio (kg/kg).
d Molecular weight (kg/kmol).
e Mole fraction (Xi ) (%).

Table 2
Summary of all test conditions
Fuel type Data d Ydil Q̇fuel φ Q̇hx Tair Tfurnace τG
points (mm) (%) (kW) (%) (◦ C) (◦ C) (s)
NG 145 2.0 0 7.6–20.4 0.67–0.98 17–51 20–780 768–1368 2.8–11.4
2 3.0 0 15.6 0.80 18 450 1295 3.9
NG/CO2 8 3.0 38–76 15.6 0.80 16–34 454 1053–1309 1.8–3.1
NG/N2 3 2.0 41–70 7.9–10.5 0.80 25–29 23–445 951–1129 3.7–4.5
5 3.0 33–71 15.6 0.80 18 450 1285 1.6–3.0
LPG 1 3.0 0 15.0 0.80 21 450 1290 3.9
LPG/CO2 10 2.0 31–71 9.9–14.4 0.75–0.81 24–41 458 1031–1261 1.3–3.5
4 3.0 50–74 15.0 0.80 20 450 1268–1306 1.1–2.0
LPG/N2 9 2.0 19–65 14.7 0.80 18–46 447 1046–1298 1.1–3.0
4 3.0 45–70 15.0 0.79 20 450 1285–1326 0.9–1.8

The baseline case of this project was defined as the the sudden decrease in furnace reference temperature
2.0-mm burner with a 15-kW thermal input (Q̇fuel ) of and by the fourfold drop in NOx emissions. The dif-
natural gas, an equivalence ratio of 0.80, and 450 ◦ C ferences seen in the levels of O2 , and consequently
air preheat (Tair ). For this condition the inlet jet ve- of CO2 , during the transition to the MILD regime
locities for air and fuel are U = 29.1 m/s and u = were caused by a slight increase in pressure drop
17.3 m/s, respectively. The reference position for the in the fuel supply system. However, it is clear that
heat exchanger was position A3 (centerline at z = these changes did not significantly influence perfor-
292.5 mm) with 25% insertion. mance. Fine-tuning of the fuel flow rate was typically
The furnace was started from a cold state with a needed to compensate for small pressure variations
nonpremixed flame stabilized by a bluff body. When in the supply line. For excess air levels of 25% or
the temperature at the lower corner of the furnace, more (φ  0.8), CO and UHC emissions were be-
which was identified as safety TC in Fig. 1, exceeded low the detection limit. From the time the furnace
autoignition (≈800 ◦ C) and the exhaust temperature was switched to the MILD regime, about 1.5 h was
exceeded a set threshold (≈600 ◦ C), the furnace was needed to reach steady-state MILD combustion con-
switched to the MILD combustion mode. ditions. Under steady-state conditions, the tempera-
ture difference between the top and the bottom sec-
tions of the furnace was around 110 ◦ C for the base-
3. Results line case. The low temperature gradient throughout
the furnace is an important characteristic of this com-
3.1. Temperature and pollutant emissions during bustion regime. The performance characteristics of
transition the furnace/burner have been discussed previously in
more detail [29]. Together they demonstrate that the
Fig. 3 shows the temperature and emissions mon- present furnace/burner configuration is able to sus-
itored as a function of time for the baseline case. The tain stable MILD combustion, while maintaining high
dashed line represents the time when MILD com- modulation ratios for various thermal input, heat ex-
bustion was activated, the bluff-body was retracted, traction, and air preheat conditions. The investigation
and no flame front was visible [29]. The successful also identified cases in which the MILD regime was
transition between the two regimes is evidenced by established without air preheat.

Please cite this article in press as: G.G. Szegö et al., Scaling of NOx emissions from a laboratory-scale mild combustion
furnace, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.02.001
ARTICLE IN PRESS CNF:6965
JID:CNF AID:6965 /FLA [mNY4a; v 1.89; Prn:5/03/2008; 14:39] P.6 (1-15)
6 G.G. Szegö et al. / Combustion and Flame ••• (••••) •••–•••

Fig. 3. Temperature and emissions recorded during the tran- Fig. 4. Measured temperature distribution (◦ C) for the base-
sition from conventional to MILD combustion for the base- line case (d = 2.0 mm, Q̇fuel = 15 kW NG, φ = 0.80,
line operating conditions (d = 2.0 mm, Q̇fuel = 15 kW NG, Tair = 450 ◦ C, 25% insertion at position A3) at plane y = 0.
φ = 0.80, Tair = 450 ◦ C, 25% insertion at position A3). The arrows represent the incoming air (solid blue) and fuel
(solid green) streams and the outgoing exhaust (dotted red)
3.2. Thermal field measurements stream. The dashed line shows the heat exchanger (HX) lo-
cation and insertion length. (For interpretation of the refer-
Measurement of the thermal field is a necessary ences to color in this figure legend, the reader is referred to
step in identifying and analyzing potential regions of the web version of this article.)
pollutant formation and/or destruction. Fig. 4 presents
the measured temperature distribution for the base- temperature measurements presented in Fig. 3. The
line case in the x–z plane through the furnace cen- uniformity suggests that a nonadiabatic global fur-
terline. The solid blue arrow represents the incoming nace temperature can reliably represent the reaction
air stream (x = 0 mm), the solid green arrows the zone temperature. Nonetheless, the most representa-
fuel streams (x = −110 mm and x = 110 mm), and tive location for such a measurement will depend on
the dotted red arrows the outgoing exhaust stream the heat exchanger location. The region in the vicinity
(x = −55 mm and x = 55 mm). The dashed line of the heat exchanger is about 100 ◦ C cooler than that
shows the heat exchanger location (window A3) and on the opposite side of the furnace. From the knowl-
insertion length. The temperature contours were gen- edge of heat extracted (Q̇hx ) and material properties,
erated using a Kriging-type interpolation algorithm the average heat exchanger surface temperature (T̄hx )
[30] from the original measurement grid, which con- was calculated assuming one-dimensional conduction
sisted of 3 × 4 × 6 points in each Cartesian direction. to be T̄hx ≈ 70 ◦ C. Hence, local temperatures around
A nonsymmetrical temperature field is noticeable the heat exchanger are expected to be even lower than
inside the furnace and is attributed to the presence 1175 ◦ C. This will result in local extinction and par-
of the heat exchanger inserted 25% through win- tial oxidation near to the cold surface.
dow A3. The highest temperature region was mea-
sured to be 1300 ◦ C in the region 80 < x < 100 mm 3.3. Effect of fuel dilution
and 200 < z < 300 mm. The temperature differen-
tial is only about 100 ◦ C within about two-thirds of A high degree of dilution of the reactants is one of
the furnace (z  180 mm). This is consistent with the the necessary conditions for establishing MILD com-

Please cite this article in press as: G.G. Szegö et al., Scaling of NOx emissions from a laboratory-scale mild combustion
furnace, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.02.001
ARTICLE IN PRESS CNF:6965
JID:CNF AID:6965 /FLA [mNY4a; v 1.89; Prn:5/03/2008; 14:39] P.7 (1-15)
G.G. Szegö et al. / Combustion and Flame ••• (••••) •••–••• 7

slightly different temperatures in each case, despite


the nominal thermal input being held constant. For
the same diluent mass fraction, the difference between
each operating condition is at most 70 ◦ C, but can be
as low as 5 ◦ C. This small difference does not explain
the variations observed in NOx emissions. Hence, it
is evident that fuel type, diluent, and burner size have
a significant effect on NOx emissions. Each of these
effects is examined in turn below and discussed in de-
tail later.
First considered are the cases in which only the
fuel type was varied, for instance LPG/CO2 (3 mm)
versus NG/CO2 (3 mm). The general trend observed
in the graph is that NOx emissions from LPG combus-
tion are higher than those from NG. Dally et al. [7] re-
ported similar trends for diluted methane and propane
flames.
Next examined are the cases in which the diluent
type is varied. Again in agreement with the results
reported by Dally et al. [7], Fig. 5 shows that fuel
dilution with CO2 (open symbols) produces lower
NOx emission levels than does the equivalent fuel di-
lution with N2 (solid symbols) for the same burner
size. Further, the CO2 dilution causes a decrease in
NOx emissions of up to 48% over the operating range,
whereas the N2 dilution causes only about 2–10%
reduction, even though the N2 content inside the fur-
Fig. 5. Effect of fuel dilution on temperature and NOx emis-
nace increases up to 19% for the highest diluent mass
sions with fuel type and burner size as parameters for the
baseline operating conditions (Q̇fuel = 15 kW, φ = 0.80, fraction cases (Ydil > 68%). The ratio of the specific
Tair = 450 ◦ C, 25% insertion at position A3). Solid sym- heats of CO2 to that of N2 is 1.076 at 1280 ◦ C. Since
bols: N2 dilution, open symbols: CO2 dilution, solid lines the furnace temperature is virtually constant, the dif-
with small symbols: d = 2.0 mm, and dashed lines with ferences in specific heats and gas radiation character-
large symbols: d = 3.0 mm. Red colored lines and symbols istics are deduced to be unable to explain the different
represent PSR calculations for methane (GRI 3.0). (For in- behaviors.
terpretation of the references to color in this figure legend, Next assessed are the results from the PSR model
the reader is referred to the web version of this article.) for the combustion of NG. The PSR calculations were
carried out using the steady-state solver from the AU-
bustion. An interesting approach to increasing vitia- RORA code of the Chemkin 4.1 package with the
tion and delaying the reaction between air and fuel GRI 3.0 [31] detailed kinetic mechanism and are rep-
is the addition of an inert gas to the fuel stream. resented in Fig. 5 by the red colored lines and sym-
Fig. 5 shows the effects of fuel dilution on temper- bols. The combustion chamber, which has a volume
ature and NOx emissions for two different fuels and of 0.0459 m3 , was modeled as a nonadiabatic reac-
burner sizes under the baseline operating conditions. tor at atmospheric pressure. The energy equation was
It also shows predictions from a perfectly stirred re- solved for a fixed heat loss of 5.5 kW, which is con-
actor (PSR) model for the NG cases. The fuel stream sistent with the energy balance shown in a previous
was diluted using either N2 or CO2 and the corre- work [29]. Four inlet streams were considered: an air
sponding diluent mass fraction (Ydil ) is displayed in stream with a mass flow rate of 6.243 g/s at 450 ◦ C,
the abscissa. The fuel and the air mass flow rates were a fuel stream with a mass flow rate of 0.305 g/s at
kept constant while the mass flow rate of diluent was 20 ◦ C, a diluent stream with mass flow rates varying
varied. according to the NG fuel dilution experiments for the
It is clear from Fig. 5 that, although dilution can be 3.0 mm burner, and a recirculation stream with 10%
expected to cool the reaction zone, this effect is small. of the total mass flow rate of reactants, 0.65 g/s, at
For each case, up to 76% fuel dilution causes a reduc- 1000 ◦ C. The latter was considered to be a mixture
tion in furnace temperature of only 2–4%, while the of inert gases (N2 /H2 O/CO2 = 0.771/0.150/0.079 by
decrease in NOx emissions can be as large as 48%. volume) and was included to represent the diluted
The figure also shows that the furnace operates at conditions of the MCF. The predictions from the

Please cite this article in press as: G.G. Szegö et al., Scaling of NOx emissions from a laboratory-scale mild combustion
furnace, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.02.001
ARTICLE IN PRESS CNF:6965
JID:CNF AID:6965 /FLA [mNY4a; v 1.89; Prn:5/03/2008; 14:39] P.8 (1-15)
8 G.G. Szegö et al. / Combustion and Flame ••• (••••) •••–•••

Fig. 6. Relationship between NOx emissions and furnace Fig. 7. Relationship between NOx emissions and exhaust
temperature, Tfurnace , for all tested operating conditions. temperature, Texhaust , for all tested operating conditions.
The solid line represents a linear curve fit for all data. The ex- The experimental data points are grouped according to fuel
perimental data points are sorted according to fuel type, dilu- input power and fuel type. No distinction is made with re-
ent, and burner size. Diamonds: NG, circles: LPG, thin line spect to fuel dilution and burner size. The thick solid black
symbols: no fuel dilution, solid symbols: N2 dilution, thick line represents a linear curve fit for all data. The other thin
line symbols: CO2 dilution, small symbols: d = 2.0 mm, and colored lines are linear curve fits for each fuel input power
large symbols: d = 3.0 mm. bin. Open symbols: all cases burning NG, and solid symbols:
all cases burning LPG.

PSR model show the same general trend observed in fuel type, diluent, and burner size. The solid line rep-
the measurements of decreasing gas temperature and resents a linear fit to all of the data. Although there
NOx emission levels with fuel dilution. However, the is a general trend of increasing NOx emissions with
maximum predicted NOx emissions are only 6 ppm furnace temperature, the wide scatter and low corre-
for both diluents. A sensitivity analysis (not shown) lation coefficient (R 2 = 0.50) show that the depen-
showed that the recirculation stream has little effect dence is weak. It is worth highlighting that very low
on NOx formation at 1280 ◦ C. It also revealed that the NOx levels, <10 ppm, are achieved for furnace tem-
reactor temperature would need to be at least 200 ◦ C peratures below 900 ◦ C. However, at temperatures
higher to reproduce the NOx emission levels observed around 1400 ◦ C, NOx emissions exceeding 60 ppm
in the experiments, which is when the thermal-NO were measured in the MILD regime.
mechanism starts to dominate. Fig. 7 shows NOx emissions data plotted against
Finally, we assess the cases in which the burner di- the exhaust temperature, which, while lower than the
ameter is varied. It is noticeable from Fig. 5 that NOx furnace temperature, avoids the bias in some data due
emissions increase when the burner size is increased. to the location of the cooling loop relative to the ther-
For a constant diluent mass fraction and thermal input mocouple. It can be seen that there is a stronger in-
from the fuel, an increase in nozzle diameter causes fluence of temperature for all data (R 2 = 0.56) and
a decrease in the fuel jet velocity by 44% and a de- for individual fuel input ranges. Even though this
crease in the jet momentum flux by the same amount general correlation is somewhat stronger, it is suffi-
for the same diluent. A lower fuel jet momentum will ciently poor to demonstrate that other parameters are
reduce the entrainment of hot furnace gases and hence also important. This result shows that a representative
change the mixing characteristics close to the burner furnace temperature alone does not adequately scale
exit. In an attempt to confirm this trend, a third fuel NOx emissions.
nozzle of diameter 4.6 mm was also tested. However, Fig. 8 shows NOx emissions plotted against heat
stable MILD combustion could not be sustained for extraction. The heat extraction, Q̇hx , is expressed as a
the baseline operating conditions, and a flame was percentage of the total thermal input. As expected, an
visible. A discussion of the stability of the MILD increase in heat extraction, which conversely means a
regime is beyond the scope of the present paper. decrease in temperature, leads to a decrease in NOx
emissions. Because the heat extraction is only part of
3.4. Global nitrogen oxides emission correlations the net heat output, it is not surprising that the correla-
tion with heat extraction is even weaker (R 2 = 0.39)
Fig. 6 presents the NOx emissions as a function than in Fig. 7.
of the furnace reference temperature (see also Fig. 1). There is no visible correlation with fuel type, dilu-
The experimental data points are sorted according to ent, or burner size, either in Fig. 6 or in Fig. 8. No

Please cite this article in press as: G.G. Szegö et al., Scaling of NOx emissions from a laboratory-scale mild combustion
furnace, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.02.001
ARTICLE IN PRESS CNF:6965
JID:CNF AID:6965 /FLA [mNY4a; v 1.89; Prn:5/03/2008; 14:39] P.9 (1-15)
G.G. Szegö et al. / Combustion and Flame ••• (••••) •••–••• 9

Fig. 8. Relationship between NOx emissions and heat ex-


tracted, Q̇hx , for all tested operating conditions. The solid
line represents a linear curve fit for all data. The experimen-
tal data points are sorted according to fuel type, diluent, and
burner size. Diamonds: NG, circles: LPG, thin line symbols:
no fuel dilution, solid symbols: N2 dilution, thick line sym-
bols: CO2 dilution, small symbols: d = 2.0 mm, and large
symbols: d = 3.0 mm.

clear trend is found in Fig. 7 with regard to fuel type


either.

3.5. Froude number scaling

To further investigate the parameters control-


ling NOx yields, the leading-order scaling approach
of Røkke et al. [32] is explored, despite its lim-
itations [12]. Although developed for buoyancy-
controlled hydrocarbon diffusion flames, it has been
applied more broadly in the past [13,16]. Based on
a simplified finite-rate chemistry that specifically in-
cludes thermal and prompt NO formation, on the
concept of laminar flamelet structure [33] and other
several assumptions, the authors derived a correla-
tion to predict the NOx emission index, EINOx . In
their scaling of open flames, the flame volume is the Fig. 9. Leading-order NOx scaling correlation,
leading-order term, and the simple correlation was EINOx (ρfuel u/d), versus fuel jet exit Froude number,
Fre , for (a) all data points, (b) all cases burning NG,
found to be equal to EINOx (ρfuel u/d) ≈ 44 Fr0.6 e , and (c) all cases burning LPG. The experimental data
where Fre is the jet exit Froude number. Weber [16]
points are sorted according to fuel type, diluent, and
proposed an extension of Røkke’s relationship to ac- burner size. Diamonds: NG, circles: LPG, thin line sym-
count for heat extraction under furnace conditions bols: no fuel dilution, solid symbols: N2 dilution, thick
for the Scaling 400 burner series operating with con- line symbols: CO2 dilution, small symbols: d = 2.0 mm,
ventional natural gas flames. He varied the propor- and large symbols: d = 3.0 mm. Blue colored line and
tionality constant according to the levels of heat ex- symbols in part (b) represent data for a reference con-
traction, whilst conforming to the same slope of 0.6. dition (17% < Q̇hx < 30%, 12 kW < Q̇fuel < 16 kW,
For example, the scaling correlation was changed to 0.78 < φ < 0.82, 400 ◦ C < Tair < 500 ◦ C). (For interpre-
EINOx (ρfuel u/d) ≈ 23.2 Fr0.6
e for 40% heat extrac-
tation of the references to color in this figure legend, the
tion. Weber’s approach resulted in a series of parallel reader is referred to the web version of this article.)
lines.
Fig. 9 shows a log–log graph of the variation of LPG. In all figures, the experimental data points are
EINOx (ρfuel u/d) with fuel jet exit Froude number. sorted according to fuel type, diluent, and burner size.
Part (a) presents all experimental data, part (b) those The theoretical correlation of Røkke et al. [32] and
cases burning NG, and part (c) those cases burning the empirical curve fit for 40% heat extraction from

Please cite this article in press as: G.G. Szegö et al., Scaling of NOx emissions from a laboratory-scale mild combustion
furnace, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.02.001
ARTICLE IN PRESS CNF:6965
JID:CNF AID:6965 /FLA [mNY4a; v 1.89; Prn:5/03/2008; 14:39] P.10 (1-15)
10 G.G. Szegö et al. / Combustion and Flame ••• (••••) •••–•••

relation exists between EINOx and Gair /Gfuel . Other


attempts to correlate EINOx with the total momen-
tum flux (Gair + Gfuel ), momentum flux ratio of fuel
to exhaust (Gfuel /Gexhaust ), and momentum flux ra-
tio of air to exhaust (Gair /Gexhaust ) have also failed
to collapse the data.

3.6. Temperature–time NOx scaling

To further assess the role of the thermal-NO


mechanism, the methodology proposed by Turns
and Myhr [35] and Turns et al. [36] is adopted.
They developed a successful global temperature–time
Fig. 10. Relationship between the NOx emission in- NOx scaling relationship that accounted for radiation
dex, EINOx , and the momentum flux ratio of air-to-fuel, losses from open turbulent jet flames. The essence of
Gair /Gfuel , for all tested operating conditions. The experi- the temperature–time scaling is the definition of an
mental data points are sorted according to fuel type, diluent, NOx production rate, which is calculated by divid-
and burner size. Diamonds: NG, circles: LPG, thin line sym- ing an average NOx concentration by a characteristic
bols: no fuel dilution, solid symbols: N2 dilution, thick line mixing time scale. The experimental NOx concentra-
symbols: CO2 dilution, small symbols: d = 2.0 mm, and tion is determined from mass-based emission indices
large symbols: d = 3.0 mm.
and stoichiometry as

previous work by Weber [16] are also displayed. The [NOx ] [mol/m3 ]
fuel jet exit Froude number (Fre ) is the ratio of jet fρrct EINOx
momentum flux to buoyant forces and is calculated =  XNO   XNO  . (4)
from u2 /gd, where g is the gravitational acceleration. MWNO X + MWNO2 X 2
NOx NOx
From Fig. 9a, it is evident that some data depart
from the correlation for the present study (NG+LPG) Adapting the definition presented by Turns and
by almost an order of magnitude. Indeed, several data Myhr [35] for open flames to enclosed multiple-jet
sets are essentially independent of the Froude number, systems, the global residence time (τG ) is charac-
indicating that other parameters affect NOx emissions terized by the combustion chamber volume and the
and the role of buoyancy is negligible. This is consis- volumetric flow rate of hot reactants as
tent with the Froude number exceeding unity by three 4VMCF fρrct
orders of magnitude (Fre > 103 ), which suggests a τG [s] = , (5)
π d 2 unρfuel
momentum-controlled system. Even those data sub-
sets that do scale with Fre , such as the correlation where VMCF is the MCF volume, f is the fuel mixture
(NG reference) represented in blue in Fig. 9b, exhibit fraction, ρrct is the density of the mixture of reactants
a slope slightly different (∝Fr0.68
e ) from those found at furnace temperature, d is the fuel nozzle diameter,
earlier [32]. Here, ∝ is used to indicate proportional- u is the fuel inlet velocity for a single nozzle, n is
ity. Costa et al. [34] found similar departures for their the number of round fuel nozzles, and ρfuel is the fuel
momentum-controlled open jet methane flames. density at ambient temperature. The use of the furnace
Figs. 9b and 9c show that the slope for NG as the volume to represent the reaction zone is a reasonable
fuel (∝Fr0.75
e ) is significantly different from that of approximation for MILD combustion conditions and
LPG as the fuel (∝Fr0.38 e ). The dependence on fuel is consistent with the global treatment used in other
type is expected from the model of Røkke et al. [32]. studies involving furnaces [14]. The fuel mixture frac-
The effects of fuel dilution can also be observed in tion is calculated from f = φ/φ + (A/F )stoic and
Fig. 9c, when the N2 dilution cases (∝Fr0.41 e ) and implicitly incorporates the air mass flow rate. Here,
the CO2 dilution cases (∝Fr0.30e ) are separately cor- (A/F )stoic is the value of stoichiometric air-to-fuel
related. ratio. The denominator in Eq. (5) is directly related to
Fig. 10 shows the variation of the NOx emission the fuel mass flow rate, which in turn determines the
index, EINOx , with the momentum flux ratio of air, fuel input power, also referred to as the firing rate.
Gair , to fuel, Gfuel . Once again, the experimental A convective time scale, d/u, is commonly used as
data points are sorted according to fuel type, diluent, the scaling parameter in momentum-controlled open
and burner size. Here Gair /Gfuel = nṁair U/ṁfuel u, jet flames. For those flames, d/u is directly propor-
where ṁair is the air mass flow rate and ṁfuel is the tional to residence time because it controls flame vol-
fuel mass flow rate. It is evident that no obvious cor- ume. The convective time scale strongly influences

Please cite this article in press as: G.G. Szegö et al., Scaling of NOx emissions from a laboratory-scale mild combustion
furnace, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.02.001
ARTICLE IN PRESS CNF:6965
JID:CNF AID:6965 /FLA [mNY4a; v 1.89; Prn:5/03/2008; 14:39] P.11 (1-15)
G.G. Szegö et al. / Combustion and Flame ••• (••••) •••–••• 11

represents the data from the present study regressed


in the same form as proposed by Turns and Myhr [35]
and Turns et al. [36], for a residence time of 4 s. Here,
[NOx ]/τG is the characteristic NOx production rate
and the regression coefficients A, B, and C are shown
in Table 3.
Several observations arise from the results shown
in Fig. 11. The use of the global residence time to
determine NOx production rates provides a general
collapse of the experimental data. However, there is
still considerable scatter in the data, which is slightly
masked by the logarithmic scales on the axis. This
scatter around a general correlation is also evident
in the data of other researchers in the MILD regime.
Fig. 11. NOx production rates calculated with the global res- For example, the data from Plessing et al. [41] and
idence time, τG , as a function of the reciprocal of the furnace Dally et al. [7] lie well within an acceptable error
temperature. The experimental data points are grouped ac-
range. Interestingly, their measurements were car-
cording to global residence time. +, CH4 [42]; ∗, NG [5];
a, CH4 [41]; ⊕, NG [17]; ×, NG [25]; ⊗, CH4 [38,40]; ∅,
ried out in a furnace with dimensions and operating
LPG [39]; ", CH4 and C3 H8 [7]. conditions similar to those for the present facility,
but with a completely different burner design, the
FLOX® burner. On the other hand, all other data ex-
the flame radiant fraction, and thus also the tempera-
cept for those of Özdemir and Peters [38,40], the
ture and NOx emissions [35,37]. In contrast, when the
rest of the data were measured in furnaces of differ-
flame volume is fixed and where exit strain does not
ent sizes that were fitted with unique burner arrange-
control flame stabilization, d/u and τG are no longer
ments. The temperature–time scaling demonstrates a
directly proportional. For this reason, d/u is not a
relevant parameter in enclosed systems, and does not dependence of NOx production rates on furnace tem-
collapse the NOx data. perature considerably weaker than that predicted by
Fig. 11 shows an Arrhenius-type plot of NOx the thermal NOx route alone. This trend applies for
production rates calculated with the global residence highly diluted furnace environments regardless of the
time, τG , as a function of the reciprocal of the fur- furnace/burner geometry. Importantly, although the
nace temperature. The experimental data points are magnitude of [NOx ]/τG depends on the combustion
grouped according to global residence time. The var- chamber volume, the slope is independent of this de-
ious blue symbols refer to NOx production rates esti- finition. Sobiesiak et al. [17] also found a departure
mated from the work of various other researchers [5, from the thermal-NO mechanism based on a well-
7,17,25,38–42]. Three different lines are also shown stirred reactor model, especially at low characteristic
in Fig. 11. The dashed line represents the theoretical temperatures.
NO production rate via the Zel’dovich thermal-NO The slope of the present data is also lower than that
mechanism. The maximum NO formation rates were for the open methane, ethylene, and propane flames of
calculated for φ = 0.80 considering a quasi-steady- Turns et al. [36]. This shows that the dependence of
state approximation for N atoms, assuming equilib- NOx production rate on temperature is stronger for
rium concentrations for O atoms, and applying the open turbulent jet flames, than for the present MILD
forward reaction coefficient, kf1 , for the rate-limiting combustion conditions. This finding is further sup-
step in the thermal-NO formation process, O + N2 → ported by the much larger value of the temperature
NO + N. The value for kf1 was computed from the coefficient, C, presented in Table 3 for open flame
expression recommended by Davidson and Hanson conditions. Nevertheless, despite this weaker depen-
[43], kf1 [m3 s/mol] = 1.95 × 108 exp(−38,660/T ). dence, it appears that the temperature–time scaling
The concentrations for the O and N atoms were calcu- correlation provides the single best description of
lated from the NASA chemical equilibrium code [44]. NOx formation in the MILD regime.
The dash-dotted line represents the two-parameter re- To evaluate the quality of the regression fit for the
gression fit present study, the percentage difference between the
  regression predictions of NOx production rates and
[NOx ] C
ln = A + B ln(τG ) + (6) the experiments is presented in Fig. 12. The analysis
τG T is performed for the complete data set using a number
for hydrocarbon flames for a residence time of 100 ms of alternative parameters substituted in Eq. (6) (cf. Ta-
extrapolated for a lower temperature range from pre- ble 3). From top to bottom in Fig. 12, the regression fit
vious work by Turns et al. [36]. Finally, the solid line assesses, respectively, furnace temperature and global

Please cite this article in press as: G.G. Szegö et al., Scaling of NOx emissions from a laboratory-scale mild combustion
furnace, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.02.001
ARTICLE IN PRESS CNF:6965
JID:CNF AID:6965 /FLA [mNY4a; v 1.89; Prn:5/03/2008; 14:39] P.12 (1-15)
12 G.G. Szegö et al. / Combustion and Flame ••• (••••) •••–•••

Table 3
Regression coefficients for Eq. (6)
Database Parameters A B C Temperature range
(ln(mol/cm3 s)) (1/s) (K) (K)
Turns et al. [36] Tflame , τG 1.1146 −0.7410 −16,347 1900–2350
Tfurnace , τG −6.141 −0.387 −3,634 1050–1650
Present study τG −8.106 −0.733 0 N/A
Tehxaust −2.249 0 −6,681 760–1070

levels of up to 400%, an almost identical fit is found


when only residence time is used, 77% of the data
are within an error band of ±50% (or 87% within
an error band ±100%). A slightly better fit is found
when the regression is based on the exhaust tempera-
ture alone, 85% of the data are within an error band
of ±50% (or 97% within an error band of ±100%).
The higher quality of the (Texhaust ) regression than of
the (Tfurnace , τG ) regression is consistent with the ex-
haust temperature being less susceptible to bias due to
the proximity to a heat sink, as previously noted. The
largest errors were found in the predictions of CO2
dilution cases. However, excluding those points did
not improve the regression significantly. Likewise, a
regression using exhaust temperature and global resi-
dence time (Texhaust , τG ) was not significantly differ-
ent.

4. Discussion

The Zel’dovich thermal mechanism is usually the


major contributor to NOx formation in most high
temperature (T > 1500 ◦ C) combustion systems. As
already seen in Figs. 4 and 6, the temperatures in
the present furnace are considerably lower than this
(T < 1360 ◦ C) and its distribution is quite uniform.
Because of the low fluctuations of MILD combustion
[45], it is unlikely that instantaneous temperatures
will significantly exceed the measured mean values.
Hence, consistent with the well-established under-
standing of the operation of MILD combustion sys-
tems, NOx formation via the thermal-NO mechanism
is inhibited relative to conventional combustion [42].
Thermal NOx not only is suppressed, but also does
not control NOx emissions, as was previously noted
Fig. 12. Percentage difference between the regression pre- in Fig. 11. The results presented in Figs. 7 and 8 re-
dictions (cf. Table 3) of NOx production rates and the exper- inforce that the thermal-NO route is not dominant.
iments for all 191 data points. Therefore, Fenimore’s prompt-NO [46] and/or N2 O-
intermediate [47] mechanisms are also expected to be
important.
residence time (Tfurnace , τG ), global residence time The fuel dilution results from Fig. 5 suggest that
(τG ) only, and exhaust temperature (Texhaust ) only as NOx formation may be dependent on the ratio of
parameters. For the (Tfurnace , τG ) case, 77% of the carbon to hydrogen atoms (C/H) in the fuel mole-
data are within an error band of ±50% (or 92% within cule. Since the temperatures are so similar for the
an error band of ±100%). Although the errors reach dilution cases, the difference cannot be attributed to

Please cite this article in press as: G.G. Szegö et al., Scaling of NOx emissions from a laboratory-scale mild combustion
furnace, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.02.001
ARTICLE IN PRESS CNF:6965
JID:CNF AID:6965 /FLA [mNY4a; v 1.89; Prn:5/03/2008; 14:39] P.13 (1-15)
G.G. Szegö et al. / Combustion and Flame ••• (••••) •••–••• 13

the thermal-NO route. Rather, the fuel type depen- differences between CO2 and N2 dilution. However,
dence is further evidence that the prompt-NO mech- the results from Fig. 10 demonstrate that, although
anism is significant. The dependence on fuel type mixing has some influence, it does not control global
has also been observed in Fig. 9. The relevance of NOx yields. Therefore other effects, such as those due
the prompt-NO pathway in the MILD regime has to chemical kinetics, also influence the NOx forma-
been speculated on by a few researchers [5,17,25, tion in this furnace.
48]. In the predictions of Mancini et al. [48], only The PSR calculations from Fig. 5 indicate that the
5% of global NOx emissions were calculated to be assumption of a well-stirred reactor (Da 1) used
of prompt-NO origin. On the other hand, it has been by some researchers [52,53] does not appear to be
observed that the N2 O-intermediate mechanism can reasonable in the present MCF system. This is con-
represent up to 65% of the total NOx formation in sistent with the recent modeling study of Mancini et
lean-premixed methane combustion in a jet-stirred re- al. [54]. They also found that the combustion chamber
actor at atmospheric pressure in a temperature range cannot be considered simply as a well-stirred reac-
from 1140 to 1570 ◦ C [49]. Further assessments are tor. They proposed a reactor network model, which
required to quantify the contributions of each path- consisted of 22 PSRs, to predict NOx emissions from
way to the overall NOx emissions. a semi-industrial-scale furnace. Their results empha-
Several parameters seem to alter NOx formation sized the importance of the mixing characteristics
in MILD combustion conditions and hence add to the close to the burner exit. This implies that the NOx for-
complexity of the scaling problem. The previous sec- mation process is not controlled by chemical kinetics.
tions have outlined the influence of heat extraction, air In summary, the analysis described above high-
preheat, oxygen concentration, dilution, and fuel type lights the fact that neither momentum nor kinetics
on global NOx emissions. While all these parameters has a controlling influence on NOx emissions, but
are not independent from each other, the combined that rather both are important. This suggests that NOx
effect can be ultimately associated with a character- formation in the present furnace may be within the
istic nonadiabatic temperature and residence time, as regime where mixing time scales are comparable to
shown in Fig. 11. However, the fairly poor fit found chemical time scales, and hence Damköhler numbers
in the regression analysis presented in Fig. 12 demon- are of order unity, Da = O(1). Therefore, in order to
strates that temperature and residence time influence, accurately predict NOx emissions from MILD com-
but do not control, the scaling of NOx . It is not pos- bustion systems, any model will need to include the
sible to isolate one or two controlling parameters, be- effects of both mixing and chemistry.
cause many parameters seem to have an influence of
comparable magnitude.
Although the Froude number has been used as a 5. Summary and conclusions
scaling parameter in furnace conditions with conven-
tional flames [16], Fig. 9 suggests that it is not the Global measurements of temperature and exhaust
most relevant parameter to scale NOx emissions from gas emissions obtained for 191 conditions in a 20-kW
furnaces operating in the MILD regime. This is be- MILD combustion furnace operating with a parallel
cause buoyancy forces are particularly weak in an jet burner system have been presented. The main find-
environment with low temperature gradients. Never- ings of the analysis are as follows:
theless, some dependence on Fre is evident, which
implies that fuel jet momentum is still an important 1. The presence of a water-cooled heat exchanger
parameter. The data for the two different burner sizes dominated the temperature gradients and caused
shown in Fig. 5 give further evidence of the influ- a nonsymmetrical temperature differential of
ence of momentum. A lower fuel jet momentum will about 100 ◦ C across the furnace.
lead to a change in the jet trajectory, much like that in 2. Overall NOx emission values as low as 5 ppmv
the strong-jet/weak-jet model of Grandmaison et al. dry at 3% O2 and as high as 66 ppmv dry at 3%
[50], and hence reduce the amount of dilution with O2 were recorded in the MILD regime for fur-
hot combustion products prior to mixing with the ox- nace temperatures of up to 1370 ◦ C.
idant stream. These changes will affect local species 3. Fuel dilution of up to 76% by mass, with either
concentrations and temperatures in the reaction zone, CO2 or N2 , reduced NOx emissions by up to
altering the ratio between mixing and reaction times, 48% and 10%, respectively. These levels of di-
i.e., the local Damköhler number (Da = τmix /τchem ). lution did not affect furnace temperatures signifi-
A low local Damköhler number can favor nonequilib- cantly, with a maximum reduction of only 56 ◦ C.
rium effects, such as superequilibrium O atoms [51]. The fuel type, diluent type, and fuel jet mo-
Both entrainment levels of hot combustion products mentum were all found to influence global NOx
and nonequilibrium chemistry may account for the yields. However, neither mixing nor chemistry

Please cite this article in press as: G.G. Szegö et al., Scaling of NOx emissions from a laboratory-scale mild combustion
furnace, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.02.001
ARTICLE IN PRESS CNF:6965
JID:CNF AID:6965 /FLA [mNY4a; v 1.89; Prn:5/03/2008; 14:39] P.14 (1-15)
14 G.G. Szegö et al. / Combustion and Flame ••• (••••) •••–•••

effects were found to be dominant, suggesting [4] A. Cavaliere, M. De Joannon, Prog. Energy Combust.
that the characteristic Damköhler number spans, Sci. 30 (2004) 329–366.
or is on the order of, unity. The failure of PSR [5] M. Katsuki, T. Hasegawa, Proc. Combust. Inst. 27
calculations to predict the NOx emissions further (1998) 3135–3146.
[6] H. Tsuji, M. Morita, A.K. Gupta, M. Katsuki, K. Kishi-
supports this hypothesis, as does the finding that
moto, T. Hasegawa, High Temperature Air Combus-
the Froude number scaling revealed a limited de- tion: From Energy Conservation to Pollution Reduc-
pendence of NOx emissions on jet momentum. tion, CRC Press, 2003.
4. A global NOx production rate, calculated based [7] B.B. Dally, E. Riesmeier, N. Peters, Combust. Flame
on a global residence time and a characteris- 137 (2004) 418–431.
tic furnace temperature, was found to give the [8] M. Flamme, Appl. Therm. Eng. 24 (2004) 1551–1559.
best scaling of the data across all conditions as- [9] P.R. Medwell, P.A.M. Kalt, B.B. Dally, Combust.
sessed. However, the dependence on temperature Flame 148 (2007) 48–61.
was found to be much weaker than that expected [10] H. Zhang, G. Yue, J. Lu, Z. Jia, J. Mao, T. Fujimori, T.
Suko, T. Kiga, Proc. Combust. Inst. 31 (2007) 2779–
based on the thermal-NO mechanism alone. This
2785.
trend also applied to other MILD combustion [11] R. Weber, J.P. Smart, W. Van der Kamp, Proc. Com-
systems, irrespective of the furnace/burner geom- bust. Inst. 30 (2004) 2623–2629.
etry. Furthermore, the regression analysis found [12] S.R. Turns, Prog. Energy Combust. Sci. 21 (1995) 361–
that neither temperature nor global residence 385.
time provides a sufficiently strong correlation to [13] A.D. Al-Fawaz, L.M. Dearden, J.T. Hedley, M. Mis-
be considered dominant. The leading-order ap- saghi, M. Pourkashanian, A. Williams, L.T. Yap, Proc.
proach showed that, although the jet exit Froude Combust. Inst. 25 (1994) 1027–1034.
[14] T.-C.A. Hsieh, W.J.A. Dahm, J.F. Driscoll, Combust.
number has been used in the past for enclosed
Flame 114 (1998) 54–80.
conventional flames, it is not the appropriate pa-
[15] A. Sayre, N. Lallemant, J. Dugue, R. Weber, Proc.
rameter for NOx scaling in the MILD regime. Combust. Inst. 25 (1994) 235–242.
This is consistent with the weak role of buoyancy [16] R. Weber, Proc. Combust. Inst. 26 (1996) 3343–3354.
in a furnace with such low temperature gradients. [17] A. Sobiesiak, S. Rahbar, H.A. Becker, Combust.
5. It is clear that no single NOx production mech- Flame 115 (1998) 93–125.
anism is dominant in the present MILD com- [18] J. Newby, B. Cain, T. Robertson, in: Proceedings of
bustion conditions. The nondominant role of the the 2nd International Seminar on High Temperature
Zel’dovich thermal-NO mechanism is consistent Combustion in Industrial Furnaces, Stockholm, Swe-
den, 2000.
with the moderate furnace temperatures (800–
[19] M. Nishimura, T. Suzuki, R. Nakanishi, R. Kitamura,
1400 ◦ C). Hence, it is deduced that the prompt- Energy Convers. Manage. 38 (1997) 1353–1363.
NO and/or N2 O-intermediate pathways are of [20] F.K. Besik, S. Rahbar, H.A. Becker, A. Sobiesiak, U.S.
significance comparable to that of the thermal- Patent 5,772,421 (June 30, 1998).
NO pathway. [21] B.E. Cain, T.F. Robertson, J.N. Newby, U.S. Patent
6,638,061 (October 28, 2003).
[22] T. Suzuki, K. Morimoto, U.S. Patent 4,439,137 (March
Acknowledgments 27, 1984).
[23] G.G. Szegö, B.B. Dally, G.J. Nathan, F.C. Christo,
in: Australian Symposium on Combustion and the 8th
The authors thank Graham Kelly and Bob Dyer for Australian Flame Days, Melbourne, Australia, 8–9 De-
their technical support in all stages of this project. The cember 2003.
financial support of the Australian Research Council [24] I. Nakamachi, K. Yasuzawa, T. Miyahara, T. Nagata,
(ARC) and the School of Mechanical Engineering of U.S. Patent 4,945,841 (August 7, 1990).
the University of Adelaide is also gratefully acknowl- [25] R. Weber, A.L. Verlaan, S. Orsino, N. Lallemant, J.
edged. Inst. Energy 72 (1999) 77–83.
[26] R. Siegel, J.R. Howell, Thermal Radiation Heat Trans-
fer, fourth ed., Taylor and Francis, New York, 2002.
[27] F.P. Incropera, D.P. DeWitt, Fundamentals of Heat and
References Mass Transfer, fourth ed., Wiley, New York, 1996.
[28] S.R. Turns, An Introduction to Combustion: Con-
[1] IEA/OECD, CO2 Emissions from Fuel Combustion: cepts and Applications, second ed., WCB/McGraw–
1971–2000, Organisation for Economic Cooperation Hill, Boston, 2000.
and Development and International Energy Agency, [29] G.G. Szegö, B.B. Dally, G.J. Nathan, F.C. Christo,
Paris, 2002. in: The Sixth Asia–Pacific Conference on Combustion,
[2] J.M. Beer, Prog. Energy Combust. Sci. 33 (2007) 107– Nagoya, Japan, 20–23 May 2007, pp. 231–234.
134. [30] J.C. Davis, Statistics and Data Analysis in Geology,
[3] S. Pacala, R. Socolow, Science 305 (2004) 968–972. Wiley, New York, 1986.

Please cite this article in press as: G.G. Szegö et al., Scaling of NOx emissions from a laboratory-scale mild combustion
furnace, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.02.001
ARTICLE IN PRESS CNF:6965
JID:CNF AID:6965 /FLA [mNY4a; v 1.89; Prn:5/03/2008; 14:39] P.15 (1-15)
G.G. Szegö et al. / Combustion and Flame ••• (••••) •••–••• 15

[31] Gas Research Institute, GRI-mech 3.0, available at [43] A.M. Dean, J.W. Bozzelli, in: W.C. Gardiner Jr.
http://www.me.berkeley.edu/gri-mech/. (Ed.), Gas-Phase Combustion Chemistry, second ed.,
[32] N.A. Røkke, J.E. Hustad, O.K. Sonju, F.A. Williams, Springer-Verlag, 2000, p. 143.
Proc. Combust. Inst. 24 (1992) 385–393. [44] S. Gordon, B.J. McBride, Report No. NASA RP-1311,
[33] N. Peters, Prog. Energy Combust. Sci. 10 (1984) 319– NASA Lewis Research Center, Cleveland, OH, 1994.
339. [45] M. Oberlack, R. Arlitt, N. Peters, Combust. Theory
Model. 4 (2000) 495–509.
[34] M. Costa, C. Parente, A. Santos, Exp. Therm. Fluid
[46] C.P. Fenimore, Proc. Combust. Inst. 13 (1970) 373–
Sci. 28 (2004) 729–734.
380.
[35] S.R. Turns, F.H. Myhr, Combust. Flame 87 (1991) 319– [47] C.T. Bowman, Proc. Combust. Inst. 24 (1992) 859–
335. 878.
[36] S.R. Turns, F.H. Myhr, R.V. Bandaru, E.R. Maund, [48] M. Mancini, R. Weber, U. Bollettini, Proc. Combust.
Combust. Flame 93 (1993) 255–269. Inst. 29 (2002) 1155–1162.
[37] G.J.R. Newbold, G.J. Nathan, D.S. Nobes, S.R. Turns, [49] R.C. Steele, P.C. Malte, D.G. Nicol, J.C. Kramlich,
Proc. Combust. Inst. 28 (2000) 481–487. Combust. Flame 100 (1995) 440–449.
[38] P.J. Coelho, N. Peters, Combust. Flame 124 (2001) [50] E.W. Grandmaison, I. Yimer, H.A. Becker, A. So-
503–518. biesiak, Combust. Flame 114 (1998) 381–396.
[51] J.F. Driscoll, R.-H. Chen, Y. Yoon, Combust. Flame 88
[39] S. Kumar, P.J. Paul, H.S. Mukunda, Proc. Combust.
(1992) 37–49.
Inst. 29 (2002) 1131–1137.
[52] M. De Joannon, A. Cavaliere, T. Faravelli, E. Ranzi,
[40] I.B. Özdemir, N. Peters, Exp. Fluids 30 (2001) 683– P. Sabia, A. Tregrossi, Proc. Combust. Inst. 30 (2005)
695. 2605–2612.
[41] T. Plessing, N. Peters, J.G. Wuenning, Proc. Combust. [53] M. De Joannon, A. Saponaro, A. Cavaliere, Proc. Com-
Inst. 27 (1998) 3197–3204. bust. Inst. 28 (2000) 1639–1645.
[42] J.A. Wünning, J.P. Wünning, Prog. Energy Combust. [54] M. Mancini, P. Schwoppe, R. Weber, S. Orsino, Com-
Sci. 23 (1997) 81–94. bust. Flame 150 (2007) 54–59.

Please cite this article in press as: G.G. Szegö et al., Scaling of NOx emissions from a laboratory-scale mild combustion
furnace, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.02.001

You might also like