You are on page 1of 23

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/261750233

CFD analysis of a syngas-fired burner for ceramic industrial roller kiln

Article in ARCHIVE Proceedings of the Institution of Mechanical Engineers Part C Journal of Mechanical Engineering Science 1989-1996 (vols 203-210) · November 2013
DOI: 10.1177/0954406213477340

CITATIONS READS

12 3,852

4 authors, including:

Marco Cavazzuti Mauro A. Corticelli


Università degli Studi di Modena e Reggio Emilia Università degli Studi di Modena e Reggio Emilia
38 PUBLICATIONS 663 CITATIONS 68 PUBLICATIONS 483 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Electric motors thermal management and optimization View project

Optimization methods View project

All content following this page was uploaded by Marco Cavazzuti on 01 December 2021.

The user has requested enhancement of the downloaded file.


CFD analysis of a syngas-fired burner for ceramic

industrial roller kiln

Marco Cavazzuti1,⇤, Mauro A. Corticelli1 ,


Antonino Nuccio2 , Bruno Zauli3
1
Dipartimento di Ingegneria “Enzo Ferrari”, Università degli Studi di Modena
e Reggio Emilia, via Vignolese 905, 41125, Modena, Italy
2
Sacmi Forni SpA, via dell’Artigianato 10, 42013, Salvaterra di Casalgrande, Italy
3
Sacmi SC, via Selice Provinciale 17/A, 40026, Imola, Italy

Abstract

Kiln burners for industrial tile production are usually fueled by methane gas.
However, the interest towards the use of coal or synthesis gases is rapidly grow-
ing, mainly due to the opening of important markets in developing countries.
The widely variable chemical composition of these fuels demands the gas burner
to be adapted case-by-case, since the firing parameters are strictly fixed, to guar-
antee the needed temperature distribution within the kiln. In this context, CFD
analysis represents a very convenient alternative to the traditional design based
on experiments. In this paper three-dimensional numerical predictions are pre-
sented for a syngas-fired burner. Three di↵erent fuels, two burner layouts, and
two burner nominal power are considered. Temperature, velocity, and oxygen
mass fraction distributions are discussed, and general design lines for low LHV
gas burners are extracted.
⇤ Corresponding author: Marco Cavazzuti, email: marco.cavazzuti@unimore.it

1
Keywords

gas burner, syngas, CFD analysis

Notation

I Unburnt fuel ratio


ṀO2 ,in Oxygen mass flow rate at the burner inlet
ṀO2 ,out Oxygen mass flow rate at the burner outlet
S Sutherland temperature
T Mixture temperature
T0 Reference temperature
µ Mixture dynamic viscosity
µ0 Reference dynamic viscosity

1 Introduction

Industrial kilns require an accurate set up and thorough control of the firing
parameters for granting the quality of the final product. Parameters like the
firing time and temperature have a huge influence on the outcome of the firing
process. Moreover, the temperature usually needs to be varied during the firing
for catalysing specific reactions at given times; also, a uniform temperature dis-
tribution is sought within the kiln section for ensuring uniform product quality.
For this reasons roller kilns are often applied in industry since they allow the
change in temperature to be achieved more e↵ectively by moving the product
through areas at di↵erent temperatures along the kiln length, more than by
warming up and cooling down the kiln periodically. Besides, they also enhance
the productivity by allowing continuous production to take place.
In particular, industrial kilns for ceramic tiles production are mainly of the
roller, gas-fired type. They can be very long (up to 300 m), and have up to
several hundred burners along their length. At the core of the firing process,
depending on the type of tiles production, they are required to be able to keep
a room temperature up to 1250 o C while the silicon carbide (SiC) burner tubes
restraining the flame are preferably kept at temperatures below 1300 o C during

2
operation.
The burner is usually fueled by methane gas. However, in developing coun-
tries, it is of interest to investigate the possibility of operating the burners using
gases coming from gasification processes. These are mixtures mainly composed
by carbon monoxide and dioxide, hydrogen, methane, and nitrogen in various
ratios. We will refer to these gases with the term syngas even though di↵erent
names can be found in literature depending on the gasification process and the
raw material employed (e.g. coal gas, producer gas, wood gas, and so on). Of
course, for achieving an efficient combustion process it is not sufficient to simply
operate a given burner using di↵erent fuels, since the burner design should be
modified in view of this. The burner head needs thus to be adapted, or re-
designed, to fulfil the firing performance of the kiln. As an additional drawback
over the employment of syngases their chemical composition is always subject
to uncertainties depending on the gasification process they come from. These
uncertainties a↵ect the firing process control and the quality of the final prod-
uct. In this context, the use of Computational Fluid Dynamics (CFD) in the
design process, including suitable combustion and radiation models, can be very
helpful for assessing the performance of the burners under a wide range of op-
erative conditions.
While in the literature many papers involving CFD analysis of kilns can be
found, they usually address furnaces for the food industry (e.g. bread mak-
ing), and the stress is always put either on improving the firing process control
and product quality, or on assessing and enhancing the energy performance by
means of heat recovery or thermal losses reduction. Although these concerns
are definitely reasonable and well-grounded, both ceramic industrial kilns, and
the use of syngases in industrial kilns seem not to have received space in the
literature at all. Yet, in this field there is still room for improvement, and small
enhancements in ceramic industrial kilns or burners design would turn out to
be very relevant in terms of energy saving, due to the large thermal powers at
stake (up to several MW).
In early papers rather elaborated lumped parameters or 1D models are found

3
for simulating the behaviour of biscuit baking ovens [1] or various type of high-,
medium-, and low-temperature heat transfer devices [2] such as bricks or glass
furnaces, cement kilns, and baking ovens.
Other papers present interesting ideas for the improvement of the energy
performance of the ovens, such as the employment of low-emissivity internal
walls [3], or the use of oxygen enriched air by using permselective membranes
for reducing fuel consumption [4]. The latter is a simple and attractive option,
especially in case a syngas is wished to be adopted as fuel. Unfortunately, both
these ideas are not applicable on large scale industrial kilns.
More recent papers all rely on the use of CFD, with special emphasis on the
food industry, mainly involving biscuits [5] and bread baking [6, 7]. Still dealing
with bread baking CFD simulations, Therdthai et al. in [8] also include a model
for estimating bread quality in terms of weight loss, crust formation, and colour,
while Boulet et al. in [9] developed a model for the transient description of the
heat transfer in a bakery pilot oven. Of course in ceramics applications things
are quite di↵erent, however the development of similar models for numerically
estimating the product quality and modeling the heat transfer would be of in-
terest.
Williamson and Wilson in [10] state the importance of radiation heat trans-
fer and air moisture content in food baking processes: the same could surely
be said for industrial ceramic kilns. They propose and numerically test a novel
radiant burner design studied for improving the uniformity of the heat flux ratio
incident on an target surface in an oven.
A di↵erent field of application in which however the temperature uniformity
is still very important is investigated by Smolka et al. in [11]. In fact, they sim-
ulate and try to improve the performance of a laboratory dry oven with forced
air circulation by proposing design changes and testing them by means of 3D
CFD.
In the frame of a wider research activity aimed at improving the perfor-
mances of roller kilns for tiles production, a predictive analysis has been car-
ried out on gas burners. Three di↵erent gases (two syngases and methane)

4
and two kind of burner tubes (having 40 mm and 50 mm outlet respectively)
are adopted in the simulations. The temperature, velocity, and oxygen mass
fraction distributions in the burner are discussed under di↵erent operating con-
ditions (12.7 kW and 38.0 kW burner nominal power). Modifications to the
burner geometry are suggested and numerically tested. The goal is to confine
the combustion within the burner body, that is, to achieve a lower ratio of un-
burnt fuel at the burner exit. This because direct radiation heat transfer from
the flame to the tiles surface would be detrimental to the final product quality
and uniformity.

2 Problem definition

2.1 Numerical domain and mesh

The model of the burner is sketched in Fig. 1, where the pre-mixing and the
burning sections are shown separately as Fig. 1(a) and Fig. 1(b) respectively.
In Fig. 1(a) air and fuel paths are also pointed out by some arrows. The mixing
part of the burner is composed by a central straight tube carrying the gas;
between this tube and the rest of the burner body flows the air entering from
a lateral inlet. The mixing takes place inside a cup-like zone where gas and air
flow-in through several holes. A narrow air passage is also present between the
cup rim and the outer burner body. The lateral duct flowing into the gas tube
at the bottom end of the burner simply serves for pressure gauge connection and
terminates in a dead-end for simulation purposes. The upper part of the burner
is simply made by a silicon carbide tube where mixing and burning processes
take place (Fig. 1(b)). The overall length of the burner body is 568 mm. The
whole burner volume was meshed using 950.000 elements; the mesh is made up
with two to four layers of hexahedral elements at the walls, and for the rest is
tetrahedrical.

5
(a) Lower part of the burner: inner tube (b) Upper part of the burner: tube restraining the
for gas passage, outer burner body for flame and mixture outlet on the top
air passage, and mixing cup on the top

Figure 1: Gas burner geometry.

2.2 Combustion model and gases

The CFD simulations were performed using Fluent. A steady-state pressure-


based implicit solver was adopted, with SIMPLE pressure-velocity coupling and
first-order upwind discretization.
Concerning the combustion model employed, an equilibrium, non-adiabatic,
non-premixed PDF combustion model [12, 13] was adopted. This is a rather
simple model in which combustion is simplified to a mixing problem. No in-
dividual species transport is solved, but two transport equations are included
instead. One for the mean mixture fraction, the other for its variance. Thermo-
chemistry calculations are pre-processed and saved in a look-up table, allowing
components concentrations to be derived from mixture fraction distribution fol-
lowing an algorithm based on the minimizazion of the Gibbs free energy [14].
The look-up table includes 60 discrete steps for the mean mixture fraction, 40
for the mean mixture fraction variance, and 80 for the mean enthalpy. Only the
8 species in Tab. 1 are tracked, thus complete one-step oxidation of the reactive
species of the fuel (namely, H2 , CH4 , and CO) is supposed. Turbulent e↵ects
are accounted for with a probability density function describing the fluctuations
of the mixture fraction due to the turbulent flow. The model is computationally
efficient and is most suitable for turbulent di↵usion flames with a rapid chemical

6
kinetic, and for problems in which the fuel and the oxidizer streams are sepa-
rate. No NOx formation model was included in the simulations at this stage.
The air and fuel inlet boundary conditions are set so that the combustion
occurs with a 10 % excess of dry air. Three di↵erent gases were taken into
consideration: two synthesis gases and methane. The reference chemical com-
positions, properties and flow rates of the three gases are summarized in Tab. 1
and in Tab. 2. Two burning powers were simulated for each geometry and fuel.
These are:

a) full power (38.0 kW),

b) one-third reduced power (12.7 kW).

These conditions are considered to cover the full operative range of a burner for
ceramic kilns.
The burner geometry investigated had specifically been designed for being
operated with syngases; in fact, being the syngas less reactive, the mixing cup
must be shaped in order to grant a stronger mixing between the fuel and the
oxidizer streams. Besides, it must also be considered that, due to the low LHV
of the syngas, the two streams will enter the mixing cup with comparable mass
flow rates. As a consequence of this, the e↵ectiveness of such a burner when
operated with methane is expected to be relatively low. Notwithstanding this,
simulations for methane were carried out for the sake of comparison. Note that
the heating value of syngases is extremely low, ranging between 7 and 13 %
towards methane, as the mass fraction of the gas e↵ectively taking part to the
combustion process (i.e. H2 , CH4 , CO) is between 0.16 and 0.42 of the total
and is mainly made of carbon monoxide.

2.3 Boundary conditions and simulations set-up

The burner walls were considered adiabatic and the mass flow rates were im-
posed at the inlets according to Tab. 2 for full power predictions, and to one-
third of the values in Tab. 2 for reduced power predictions. A pressure outlet
condition with zero static gauge pressure applies on the outlet section. The

7
Table 1: Gas and combustive agent chemical composition.

Chemical species Syngas #1 Syngas #2 Methane Dry air

Name and Atomic Stoichiometric Mass Mole Mass Mole Mass Mole Mass Mole

molecular formula weight air/fuel mass ratio fraction fraction fraction fraction fraction fraction fraction fraction

Hydrogen H2 2.02 34.290 0.014 0.158 0.009 0.111 — — — —

Methane CH4 16.04 17.235 0.014 0.020 0.022 0.035 1.000 1.000 — —

8
Carbon monoxide CO 28.01 2.468 0.389 0.325 0.136 0.126 — — — —

Carbon dioxide CO2 44.01 — 0.047 0.025 0.278 0.163 — — — —

Oxygen O2 32.00 — 0.003 0.002 — — — — 0.231 0.210

Nitrogen N2 28.01 — 0.479 0.400 0.450 0.415 — — 0.756 0.781

Water vapour H2 O 18.02 — 0.054 0.070 0.105 0.150 — — — —

Argon Ar 39.95 — — — — — — — 0.013 0.009


Table 2: Gas and combustive agent properties and mass flow rates. For normal
temperature and pressure (NTP) conditions 273.15 K and 101325 Pa are meant.

Syngas Syngas
Property and unit Methane
#1 #2
kg
Density at NTP m3 1.043 1.152 0.716

Stoichiometric air/fuel mass ratio 1.653 1.009 17.235


MJ
Lower heating value (LHV) kg 6.26 3.51 50.01
MJ
Higher heating value (HHV) kg 6.64 3.82 55.68
g
Gas mass flow rate at 38.0 kW s 6.07 10.83 0.76
g
Air mass flow rate at 38.0 kW s 11.05 12.03 14.41
g
Oxygen mass flow rate at 38.0 kW s 2.57 2.78 3.34
Nm3
Gas volume flow rate at 38.0 kW h 20.97 33.86 3.82
Nm3
Air volume flow rate at 38.0 kW h 30.77 33.52 40.13
Nm3
Oxygen volume flow rate at 38.0 kW h 6.49 7.02 8.41

k-✏ realizable turbulence model with enhanced wall functions and the Discrete
Ordinates (DO) radiation model are adopted. The walls emissivity is set equal
to 0.9 and a four angles discretization is adopted both in polar and azimuthal
directions. The specific heat of the air/gas mixture was computed through the
mixing law, while the thermal conductivity was assumed to be linearly depen-
dent on temperature, and the viscosity to follow the Sutherland law
✓ ◆ 32
T0 + S T
µ = µ0 . (1)
T +S T0

Here µ is the mixture viscosity, T its temperature, µ0 , T0 , and S are constants,


5
in detail: µ0 = 1.716 ⇥ 10 Pa s, T0 = 273.11 K, S = 110.56 K.
The choice of having adiabatic walls, even though the temperature inside a
burner is very high, is suggested by the fact that the burner body is installed
across the kiln walls, and thus, is surrounded by refractory low thermal conduc-

9
Figure 2: Mesh resolution study: temperature profile along the burner axis for
the two meshes tested.

tivity material.

2.4 Mesh resolution

A mesh resolution study was performed on a preliminary run where a 18.1 kW


syngas #1 fuelled burner running with a 15 % excess of air was tested. The mesh
used in this study, having approximately 950 thousand elements, was tested
against a mesh having roughly 7.5 million elements. Due to the complexity of
the burner geometry and, as a consequence, of the meshing procedure, the fine
mesh originates from splitting every element of the initial mesh into 8 elements.
It is found that with the fine mesh the combustion process is slightly slower due
to smaller numerical di↵usion. The temperature at the burner outlet is reduced
from 1864 K to 1829 K, the unburnt fuel ratio raised from 10.2 % to 12.6 %. The
temperature profile along the burner axis confirms this trend as shown in Fig. 2.
Even though some small di↵erence is noted between the two meshes tested, the
thinner mesh was adopted as a good trade-o↵ between accuracy and computing
time: one simulation run with the thinner mesh on a 8 processors machine took
almost 3 days to be completed.

10
3 Results

3.1 Burner simulations

Overall, twelve cases were simulated. The convergence of the simulations was
very satisfactory, with maximum residuals for the problem variables in the range
8 5
between 10 and 10 . The cost of each simulation in terms of CPU time
ranged between four to six days on a single CPU instead. Results are summa-
rized in Tab. 3. In detail, data are presented in terms of oxygen mass flow rate
at the burner outlet, average temperature and velocity of the exhaust gas. The
average pressure drop between the inlet and the outlet on both the gas and the
air sides are also given. From the oxygen mass flow rate at the burner outlet
the ratio of unburnt fuel can be estimated considering that the combustion oc-
curs with a 10 % excess of air. Under the assumption of uniform, perfect, and
complete combustion of the gas

ṀO2 , in ṀO2 , out 11 ṀO2 , out 1


I=1 = , (2)
ṀO2 , in
ṀO2 , in 10 ṀO2 , in 10
11

holds, where I is the ratio of unburnt fuel, ṀO2,in and ṀO2,out the oxygen
mass flow rate entering the burner from the inlets and exiting from the outlet,
respectively. The data show that there is a relatively high ratio of unburnt fuel
gas at the burner outlet. This is particularly true when the burner is fueled
by methane (up to 32 %). It follows that a rather low outlet temperature is
found when using this fuel, even lower than that obtained using syngas #1.
This result conforms to expectations and confirms that the gas burner design
needs to be optimized case-by-case. On the other side, syngas #2, despite the
much lower unburnt ratios, shows the lowest outlet temperature due to its poor
heating value. The better performance in terms of unburnt fuel ratio of syngas
#2 is to be ascribed to the larger fuel mass flow rate required for a given burner
nominal power (see also Tab. 2). This translates in higher fuel velocities, higher
turbulence, and better mixing in the mixing cup, thus in a faster combustion.
Figures 3 and 4 compare the temperature and the oxygen mass fraction
fields inside the burner for syngas #2, 50 mm, 38 kW, and methane, 50 mm,

11
Table 3: Burner outflow properties.

Property and unit

O2 mass Unburnt Average Average Fuel pressure Air pressure


Case Study
flow rate fuel ratio Temperature Velocity drop drop
g m
s K s Pa Pa

Syngas #1 12.7 kW 40 mm 0.1658 0.1140 1915.7 23.84 159.4 145.9

Syngas #1 12.7 kW 50 mm 0.1792 0.1313 1921.0 15.48 119.5 106.0

Syngas #1 38.0 kW 40 mm 0.5363 0.1307 1840.4 68.22 1334.5 1214.8

12
Syngas #1 38.0 kW 50 mm 0.5748 0.1473 1824.0 43.73 995.2 875.7

Syngas #2 12.7 kW 40 mm 0.1364 0.0616 1542.7 25.77 342.5 180.9

Syngas #2 12.7 kW 50 mm 0.1457 0.0726 1541.8 16.66 285.2 123.5

Syngas #2 38.0 kW 40 mm 0.4603 0.0818 1501.0 74.86 2932.4 1523.1

Syngas #2 38.0 kW 50 mm 0.4901 0.0936 1491.5 48.06 2451.6 1030.0

Methane 12.7 kW 40 mm 0.4034 0.2992 1662.3 19.48 56.0 180.2

Methane 12.7 kW 50 mm 0.4233 0.3188 1647.2 12.49 26.2 150.5

Methane 38.0 kW 40 mm 1.0675 0.2521 1713.3 59.68 488.1 1558.5

Methane 38.0 kW 50 mm 1.1286 0.2722 1684.8 37.92 219.7 1292.7


(a) syngas #2, 50 mm outlet, 38.0 kW power

(b) methane, 50 mm outlet, 38.0 kW power

Figure 3: Temperature field inside the burner: the temperature is expressed in


K.

(a) syngas #2, 50 mm outlet, 38.0 kW power

(b) methane, 50 mm outlet, 38.0 kW power

Figure 4: Oxygen mass fraction field inside the burner.

38 kW cases. Figures 5(a) and 5(b) show the temperature and oxygen mass
fraction profiles along the burner axis. For methane, due to the high air-to-
fuel ratio, a rather high oxygen ratio is found at the burner axis within a few
centimeters distance from the nozzle, at the same time, as the combustion begins
to take place, also the axial temperature gradient is much higher in the area of
the burner mixing cup. Despite this, at a higher distance from the nozzle the
combustion slows down and the oxygen is not consumed as quickly as in the
other cases (see the di↵erent low oxygen areas in Figs. 4(a) and 4(b)).
The di↵erent mass flow rates and burner nozzle diameters are the reasons

13
(a) Axial temperature (b) Axial oxygen mass fraction

(c) Axial velocity

Figure 5: Burner axial temperature, oxygen mass fraction, and velocity as a


function of the distance from the gas inlet into the mixing cup. The data for
the 40 mm cases in subfigures (a) and (b) are not shown since the di↵erences
from the corresponding 50 mm cases are unremarkable. The data for the 12.7 kW
1
cases in subfigure (c) are not shown since they scale almost perfectly to 3 of the
38.0 kW cases.

for the rather di↵erent average velocities and pressure drops found among the
twelve cases (see Tab. 3). Figures 5 and 6 show how the nozzle diameter a↵ects
the velocity field near to the nozzle itself. Away from the nozzle the velocity
field results una↵ected instead.
Fig. 7 compares the carbon dioxide mass fraction fields in syngas #1 and
syngas #2 cases. It is curious to point out how, due to the large CO2 mass
fraction initially present in syngas #2, after mixing with air and combustion,
the CO2 mass fraction of the mixture at the outlet is lower than that of the

14
(a) syngas #1, 40 mm outlet, 38.0 kW power

(b) syngas #1, 50 mm outlet, 38.0 kW power

m
Figure 6: Velocity field inside the burner: the velocity is expressed in s .

(a) syngas #1, 50 mm outlet, 38.0 kW power

(b) syngas #2, 50 mm outlet, 38.0 kW power

Figure 7: Carbon dioxide mass fraction field inside the burner.

fuel.
The asymmetry seen in the fields shown in Figs. 3 to 6 is due to the non
fully symmetrical shape of the mixing cup, and mostly to the fact that the air
inlet is located on a side of the burner body, thus promoting a non-symmetrical
air flow through the mixing cup holes.

3.2 Unburnt gas reduction

Some additional CFD analyses on the burner have been performed in order to
assess how minor and cost e↵ective changes to the burner mixing cup could in-
fluence the unburnt gas ratio at the burner nozzle. The idea is to try to improve

15
Figure 8: Holes and passages for the gas and the air flow into the mixing cup.

the performance of the burner (i.e. to reduce the unburnt gas ratio) by simply
removing or reducing in size the holes and the other passages through which
the fuel and the air flow into the mixing cup. The passages are summarized in
Fig. 8 and listed below:

• A: annular gap between the mixing cup rim and the external burner body
for the air flow,

• B: holes on the bottom floor of the mixing cup for the air flow,

• C: holes on the upper side wall of the mixing cup for the air flow,

• D: holes on the lower side wall of the mixing cup for the air flow,

• E: central hole on the bottom part of the mixing cup for the fuel flow,

• F: lateral holes on the bottom part of the mixing cup for the fuel flow.

Six additional simulations have been performed taking as a reference model


the syngas #1, 50 mm nozzle, 38.0 kW case. In each simulation one of the
items listed above has been completely closed except for case “E” in which the
hole diameter was reduced. On the basis of the results obtained, an additional
simulation was performed in which both the annular gap “A” and the central
hole “E” have been changed with respect to their original size. The results of

16
Table 4: Mass flow rates through the mixing cup passages, pressure drops in-
crement, and unburnt gas ratios for the seven additional burner configuration
tested.

Mass flow rate [%] Pressure Unburnt

Configuration Air Fuel drop [%] gas

A B C D E F Air Fuel ratio [%]

Standard 58.5 29.5 8.7 3.3 81.7 18.3 — — 14.7

A removed – 59.8 30.5 9.7 81.6 18.4 +16.4 0.9 11.8

B removed 80.7 – 13.5 5.8 81.2 18.8 +4.0 1.9 17.5

C removed 62.6 33.3 – 4.1 81.8 18.2 0.5 1.5 15.0

D removed 59.8 31.5 8.7 – 81.7 18.3 0.9 1.1 14.6

E reduced 62.1 21.8 12.1 4.0 48.6 51.4 +3.2 +56.7 7.0

F removed 53.0 33.4 9.7 3.9 100 – 3.5 +41.8 13.3

A rem. & E red. – 55.0 34.6 10.4 48.9 51.1 +44.5 +54.8 5.4

these simulations are shown in Tab. 4 in terms of mass flow rates percentage
through the sets of passages, pressure drops, and unburnt gas ratio at the burner
outlet.
From the results of the standard configuration it is clear that a large air
mass fraction (58 %) is flowing through the gap “A”. As this gap is located far
from the fuel inlets the air flowing through it is not coming into contact with
the fuel before reaching the upper part of the burner tube. For what concerns
the fuel flow, 82 % of the fuel mass flows through the central hole “E” creating
a compact jet which again has a relatively low chance to come into contact with
air, as shown in Fig. 3(a) where the core of the central fuel jet remains cold
for a while after entering the mixing cup. From these considerations, it comes
out that a stronger mixing of the two flows in the lower part of the mixing cup
would be most desirable. As a consequence, it is not strange that the solutions

17
(a) syngas #1, 50 mm outlet, 38.0 kW (b) syngas #1, 50 mm outlet, 38.0 kW
power, standard configuration power, “A” removed & “E” reduced con-
figuration

Figure 9: Turbulent kinetic energy field inside the burner mixing cup: k is
m2
expressed in s2 .

giving the better performances in terms of unburnt gas ratio reduction are the
removal of the gap “A” ( 20 %) and the reduction of the hole “E” size ( 52 %).
Putting the two changes together (last line of Tab. 4) a reduction of 63 % was
achieved. In this configuration, 55 % of the air flows from the very bottom of the
mixing cup (holes “B”), while the fuel flow is almost equally balanced between
the central hole “E” (49 %) and the lateral holes “F” (51 %), thus promoting a
better mixing of the flows and increased turbulence within the cup. The price
to pay in this case is a larger pressure drop in the burner (+55 % on the fuel
side and +45 % on the air side). This is indirectly confirmed by the comparison
between the distributions of the turbulent kinetic energy k in the standard
configuration and in the final configuration as shown in Fig. 9. In the latter
case, k increases more than three-fold.
The standard and the final configurations are also compared in terms of CO
distributions in Fig. 10, where the better performance of the latter is confirmed
due to the better mixing and lower amount of fuel through the central hole “E”.

18
(a) syngas #1, 50 mm outlet, 38.0 kW power, standard configuration

(b) syngas #1, 50 mm outlet, 38.0 kW power, “A” removed & “E” re-
duced configuration

Figure 10: Carbon monoxide mass fraction field inside the burner.

4 Conclusions

CFD analyses of an industrial burner for ceramic tiles have been performed. The
study aimed at assessing the behaviour of a properly designed burner in terms
of temperature, oxygen mass fraction, and velocity distributions using di↵erent
synthesis gases in place of methane. Di↵erent burner powers and outlet sections
were tested. It is found that the original burner geometry is not able to allow
the whole combustion process to take place inside the burner. In fact, at the
burner nozzle a relatively large amount of fuel (ranging from 6 to 32 %) remains
unreacted. Curiously, due to the larger unburnt fuel ratios, methane is found to
give lower average temperatures at the burner nozzle when compared to syngas
#1, in spite of its much higher heat power.
Additional simulations have been performed to inspect the e↵ect of small
changes in the mixing cup geometry on flow mixing. It is found that two minor
changes in the mixing cup configuration allowed the unburnt gas ratio at the
burner outlet to be reduced by 63 %.
Low LHV fuels have low stoichiometric air/fuel ratios and require much
larger mass flow rates for achieving a given burner nominal power. Leaving the
unburnt ratio influence behind, this reflects in lower gas mixture temperature
at the burner outlet and higher velocities. Of course, the velocity is even higher

19
when the smaller nozzle diameter is adopted.
The analysis has demonstrated the potential of CFD analysis in this domain,
since, in fact, not only the performance of a specific burner design can be rapidly
tested for di↵erent fuels, but also design hints can be derived with little e↵ort and
economically, towards the optimization of burner components. In this context,
numerical results demonstrate the crucial role of the mixing cup in the burning
process, since, in fact, minor modifications in its geometry can provide definite
improvements in the burner performances. Knowledge of the temperature and
velocity fields in the burner body coming from the CFD simulations, and in
particular of the mixing of the fuel and oxidizer streams, can thus lead the
designer in choosing new layouts for the mixing cup; these can go from simple
changes in the holes number and distribution, as shown in the present study, to
more complicated and time-consuming changes in shape.
It must be stressed out that for syngas burners such as those considered here,
an improvement in the burning performance is important achievement since it
translates not only into a lower fuel consumption for a given thermal power, but
also into a lower release of carbon monoxide in the atmosphere.
In the following studies it would be of interest to simulate the kiln room, or at
least part of it due to its large dimension, to see how the outlet burner conditions
reflect over the tiles floor temperature distributions since this is the real product
quality parameter tiles producers are interested in. Such an analysis would also
allow some considerations in terms of the energy balance of the whole kiln to
be drawn.
Compatibly with the cost in terms of CPU time, it would also be of interest
a thorough test of more elaborated combustion models, including models for
assessing the NOx production of the burner.

References

[1] I. Savoye, G. Trystram, A. Duquenoy, P. Brunet, and F. Marchin. Heat and


mass transfer dynamic modelling of an indirect biscuit baking tunnel-oven.

20
part I: modelling principles. J. Food Eng., 16:173–196, 1992.

[2] M. Carvalho and M. Nogueira. Improvement of energy efficiency in glass-


melting furnaces, cement kilns and baking ovens. Appl. Therm. Eng.,
17:921–933, 1997.

[3] B. M. Shaughnessy and M. Newborough. Energy performance of a low-


emissivity electrically heated oven. Appl. Therm. Eng., 20:813–830, 2000.

[4] G. Bisio, A. Bosio, and G. Rubatto. Thermodynamics applied to oxygen


enrichment of combustion air. Energy Convers. Manag., 43:2589–2600,
2002.

[5] P. S. Mirade, J. D. Daudin, F. Ducept, G. Trystram, and J. Clément. Char-


acterization and CFD modelling of air temperature and velocity profiles
in an industrial biscuit baking tunnel oven. Food Res. Int., 37:1031–1039,
2004.

[6] S. Y. Wong, W. Zhou, and J. Hua. Robustness analysis of a CFD model


to the uncertainties in its physical properties for a bread baking process.
J. Food Eng., 77:784–791, 2006.

[7] S. Y. Wong, W. Zhou, and J. Hua. Designing process controller for a


continuous bread baking process based on cfd modelling. J. Food Eng.,
81:523–534, 2007.

[8] N. Therdthai, W. Zhou, and T. Adamczak. Three-dimensional CFD mod-


elling and simulation of the temperature profiles and airflow patterns during
a continuous industrial baking process. J. Food Eng., 65:599–608, 2004.

[9] M. Boulet, B. Marcos, M. Dostie, and C. Moresoli. CFD modeling of heat


transfer and flow field in a bakery pilot oven. J. Food Eng., 97:393–402,
2010.

[10] M. E. Williamson and D. I. Wilson. Development of an improved heating


system for industrial tunnel baking ovens. J. Food Eng., 91:64–71, 2009.

21
[11] J. Smolka, A. J. Nowak, and D. Rybarz. Improved 3-D temperature unifor-
mity in a laboratory drying oven based on experimentally validated CFD
computations. J. Food Eng., 97:373–383, 2010.

[12] S.B. Pope. PDF methods for turbulent reactive flows. Prog. Energ. Com-
bust., 11:119–192, 1985.

[13] C. Dopazo. Recent developments in PDF methods. In P.A. Libby and


F.A. Williams, editors, Turbulent Reacting Flows, pages 375–474. Academic
Press, London, 1994.

[14] K.K. Kuo. Principles of combustion. John Wiley & Sons, 2nd edition,
1986.

22

View publication stats

You might also like