You are on page 1of 19

Combustion Science and Technology

ISSN: 0010-2202 (Print) 1563-521X (Online) Journal homepage: https://www.tandfonline.com/loi/gcst20

LIQUID FUELS-FIRED POROUS BURNER

S. JUGJAI & C. PONGSAI

To cite this article: S. JUGJAI & C. PONGSAI (2007) LIQUID FUELS-FIRED POROUS BURNER,
Combustion Science and Technology, 179:9, 1823-1840, DOI: 10.1080/00102200701260179
To link to this article: https://doi.org/10.1080/00102200701260179

Published online: 30 Jul 2007.

Submit your article to this journal

Article views: 275

View related articles

Citing articles: 7 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=gcst20
Combust. Sci. and Tech., 179: 1823–1840, 2007
Copyright Q Taylor & Francis Group, LLC
ISSN: 0010-2202 print/1563-521X online
DOI: 10.1080/00102200701260179

Liquid Fuels-Fired Porous Burner

S. Jugjai and C. Pongsai


Combustion and Engine Research Laboratory (CERL), Department
of Mechanical Engineering, Faculty of Engineering, King Mongkut’s
University of Technology Thonburi, Bangkok, Thailand

Abstract: This article presents an experimental study of evaporation and


combustion in porous media without atomization of the liquid fuel (kerosene).
The concept is called a liquid fuels-fired porous burner (LFPB). This concept is
based on the evaporation of the liquid fuel in an upstream porous matrix, a
subsequent mixing with the combustion air in a swirling chamber and the final
combustion in a downstream porous matrix. The evaporation is maintained by
the upstream heat transfer from the combustion zone or the downstream porous
matrix to the evaporation zone in the upstream porous matrix. The major inno-
vation of the study will be to show that the LFPB technology is able to replace
conventional spray burners. Practical applications of this new combustion system
are suggested.

Keywords: Burners; Combustion; Liquid fuels; Phase-change; Porous medium;


Radiation

INTRODUCTION

About 85% of today’s primary world energy demands rely upon fossil
fuels (coal, oil and gas), and oil plays the dominant role with a share
of about 38% (Hein, 2004). Future projections expect that this tendency
will not change substantially within the next 2 decades. Liquid fuels are
important in power generation, transportation and propulsion. Hence,
combustion of liquid fuels including liquid biofuel is expected to play a

Received 3 November 2005; accepted 30 January 2007.



Address correspondence to sumrueng.jug@kmutt.ac.th
1824 S. Jugjai and C. Pongsai

critical role so long as the combustion process is environmentally


compatible and economical.
Within the framework of the present study, a new concept of liquid
fuel combustion by inert porous burner technology is focused and
explored. In the past, combustion of liquid fuel by the porous burner tech-
nology was studied both experimentally (Kaplan and Hall, 1995; Tseng
and Howell, 1996; Takami et al., 1998; Jugjai et al., 2002; Jugjai and
Polmart, 2003; Fuse et al., 2003, 2005) and theoretically (Martynenko
et al., 1998; Kayal and Chakravarty, 2005; Park and Kaviany, 2002).
The use of a porous burner in combustion systems results in many
advantages over conventional spray burner systems. These are a heat
recirculating combustion characteristic, high combustion density, a com-
plete combustion with low emission of CO and NOx, and a relatively high
and uniform heat flux. Moreover, with specific types of porous burner
(Jugjai et al., 2002; Fuse et al., 2003, 2005; Jugjai and Polmart, 2003),
relatively low oil pressure (less than 14.7 psig) can be employed. This
can be achieved by utilizing a trickling flow of the liquid fuels through
a porous medium instead of a very fine droplet atomization, resulting
in an energy saving compared to the atomization. The porous burner
is, thus, compact in size, yielding a satisfactorily wide stable combustion
region, fuel flexible owing to enhanced evaporation, low in cost and
potentially capable of replacing conventional spray combustion. Possible
practical applications of the porous burner could be seen from several
examples, such as in an internal combustion engine (Bartsch, 2001;
Durst and Weclas, 2001; Weclas, 2001; Park and Kaviany, 2002) in a
micro gas turbine (Liedtke and Schulz, 2003) for power generation and
so on. It is worth mentioning that the porous burner technology is an
important option to replace its rival conventional spray burner tech-
nology for cleaner combustion and higher efficient energy conversion
in the future.
Even though research and development on porous burner for liquid
fuel combustion has been underway for years, problems still exist. Early
research was focused on combustion with relatively highly volatile liquid
fuels (heptane) (Kaplan and Hall, 1995; Tseng and Howell, 1996). More-
over, atomization obtained by a high-pressure injector was needed. Thus,
the porous burner is less efficient as compared to the conventional spray
burner because the full function of the porous medium has not yet been
extended for an evaporation process. Both combustion air and fuel drop-
let spray were supplied to the porous burner in the manner of an up-flow
spray porous burner. Combustion took place within the porous burner,
whereas most of the evaporation took place outside the porous burner
(Tseng and Howell, 1996) at a pre-combustion zone due to very fine
droplet spray (approximately 15–20 microns) and relatively high vola-
tility. On the other hand, an experiment of an up-flow non-spray porous
Liquid Fuels-Fired Porous Burner 1825

burner was also conducted (Fuse et al., 2003, 2005). This time evaporation
completely took place within the porous burner, but the combustion took
place outside the porous burner to its downstream end. A method allow-
ing both evaporation and combustion to simultaneously take place
within the porous burner is of special importance for its high combustion
efficiency and high radiation output for practical applications. This has
been found only in theoretical work (Martynenko et al., 1998; Kayal
and Chakravarty, 2005). In addition, there are few extensive studies in
the literature on evaporation and combustion of liquid fuels within inert
porous media.
To solve the problems with state-of-the-art technology, and to
improve the porous burner technology for liquid fuels combustion, a
new combustion concept of a liquid fuel-fired porous burner (hereafter
referred to as LFPB) under adiabatic condition at atmospheric pressure
and without a fine droplet atomization has been proposed by Jugjai
and Polmart (2003). The authors succeeded in realizing the simultaneous
phenomenon of evaporation and combustion taking place within inert
porous media through experiment. Evaporation was achieved by a
porous medium called a porous burner (hereafter referred to as PB),
whereas combustion occurred in another porous medium called a porous
emitter (hereafter referred to as PE) installed downstream of the PB.
These two porous media are separated by a small space for mixing
between fuel vapor and combustion air.
During normal operation, these two porous media are coupled by
thermal radiation for the desired continuous evaporation and combus-
tion. With this arrangement of porous media, evaporation is thus com-
pletely separated from combustion, resulting in a nearly homogeneous
mixture of fuel vapor and air to be formed in the mixing chamber fol-
lowed by a nearly homogeneous combustion afterwards. However, our
previous LFPB system (Jugjai and Polmart, 2003) was limited to only
a down-flow combustion, thus limiting its practical applications. Burner
operation in various orientations has not yet been fully understood. To
clarify the effect of the burner orientation, and to broaden the base for
future research, this experimental study was carried out.

STATE-OF-THE-ART TECHNOLOGY
OF HEAT-RECIRCULATING COMBUSTION

Heat-recirculating combustion is a promising scheme to enhance com-


bustion (Weinberg, 1971). Figure 1 shows a comparison in temperature
histories of premixed combustion in a one-dimensional adiabatic
system for the cases with and without heat-recirculation. With heat-
recirculation, reactants are preheated prior to the flame zone by heat
1826 S. Jugjai and C. Pongsai

Figure 1. Concept of heat-recirculating combustion.

transfer from burned products without mixing two streams. Despite


the same value of the final exit temperatures, there is no ceiling on
the maximum combustion temperature for the case with the heat-
recirculation depending on the amount of heat recycled. Thus, a considerable
increase in flame temperature and reaction rate can be realized-subject to
the thermal limitation of the heat exchanger and of pre-ignition of the
reactants. Therefore optimization of temperature, combustion efficiency
and emission of pollutants of this kind of combustion are important.
The present study will provide an answer to the question of what happens
when the heat-recirculating combustion technique is applied to liquid fuel
through the proposed LFPB without atomization.

EXPERIMENTAL FACILITY

Figure 2 shows an experimental setup of LFPB and Figure 3 shows its


details. The LFPB is divided into three sections: the evaporation zone
(or porous burner, PB), the combustion or mixing or swirling chamber
zone and the combustion zone (or porous emitter, PE). Fundamentally,
the LFPB is a phase-change problem in a solid matrix of PB (Kaviany,
1995), of which its downstream surface (or exit at x ¼ 75, see Figure 3)
is subjected to an intense thermal radiation emitted from flame or from
the PE for evaporation within the PB, whereas the upstream one
(x ¼ 0) is supplied by the liquid fuel without atomization. The LFPB
can produce either a surface stabilized flame (SSF) (a flame stabilized
outside and near exit of PB) or a matrix stabilized flame (MSF) (a flame
Liquid Fuels-Fired Porous Burner 1827

Figure 2. Experimental setup of LFPB.

wholly stabilized inside PE) depending on installation of the porous


emitter (PE). With the PE installed, the LFPB can yield a MSF, otherwise
it yields a SSF. For the sake of completeness, the LFPB with the PE
installed is used for describing the experimental apparatus as shown in
Figures 2 and 3.
The evaporation zone is called porous burner PB, even though it does
not contain a combustion zone as its name implies. However, it is PB that
causes evaporation followed by combustion near its exit (x ¼ 75). There-
fore, we retain the name ‘‘PB’’ so as to reflect its main contribution to
combustion via its true function as an evaporator. PB is formed by a
stack of pieces of stainless steel wire net having 60–100 mesh per inch
with an overall geometrical length of LPB ¼ 75 mm. Equivalent optical
thickness of PB, sPB is equal to 86–131, depending on mesh size of the
steel wire net. The optical thickness of the PB (also for PE) provided in
the present study as shown in Table 1 was just an estimation so as to
quantify the radiative property of the PB=PE. In the present study, the
optical thickness (s) is approximated by production of an absorption
coefficient (j) of the whole bed and a geometric length L of the bed
(s ¼ jL). j can be approximated by the production of emissivity of the
inside medium of particles and the surface area-to-volume ratios of
the bed of PB=PE, which can be calculated from their geometrical
dimensions, such as diameter of particle or wire and number density of
the particle.
1828 S. Jugjai and C. Pongsai

Figure 3. Details of LFPB.

There is a fuel injector (syringe) with its tip making direct contact
with the upstream end of the PB. Kerosene in the storage tank is pressur-
ized by nitrogen gas at relatively low pressure (about 1 atm), which is just
enough to cause the kerosene to flow to the syringe and through the PB.
Since flow, heat and mass transfer within the PB are complicated and are
governed by some important parameters, such as type of material, micro-
structure, porosity, permeability, buoyancy, capillary action and gravity
action, change in orientation of the LFPB may affect its performance. To
prove its potential for practical applications, the LFPB is equipped with a
rotating axle (not shown) perpendicular to its center line so as to allow
for variation in the burner angle h(Figure 3).
The swirling chamber is a space separating the PB and the PE with a
constant inter-distance d ¼ 50 mm. This chamber is supplied by swirling
Liquid Fuels-Fired Porous Burner 1829

Table 1. Experimental conditions

Quantity Value

Absorption coefficient of PB, jPB, m 1 1,150–1,750


Absorption coefficient of PE, jPE, m 1 13.48
Average diameter of alumina sphere of PE, dp, mm 19
Burner angle, h, degrees 0–180
Equivalence ratio, U 0.34–0.9
Heat input, CL, kW 3.5–11
Inter-distance between PB and PE, d, mm 50
Length of PB, LPB, mm 75
Length of PE, LPE, mm 210
Lower heating value of kerosene, MJ=kg 42
Number of mesh per inch of wire net of PB, mesh=inch 60–100
Optical thickness of PB, sPB ¼ jPB LPB 86–131
Optical thickness of PE, sPE ¼ jPE LPE 2.78
Pitch of wire net of PB, p, mm 0.423–0.254
Porosity of PB, ePB 0.72–0.61
Porosity of PE, ePE 0.43
Surface area-to-volume ratio of PB, m2=m3 7,420–12,370
Wire diameter of wire net of PB, dw, mm 0.15–0.125

air from three directions (see section A-A in Figure 3) for obtaining good
mixing with the fuel vapor emerging from the PB exit. Each flow direc-
tion has equal air flow rate and is directed towards the circumferential
of a small, imaginary circle at the center of the swirling chamber. Com-
bustion can partially or wholly take place within the swirling chamber,
depending on configuration of the LFPB (with or without PE).
In contrast to the PB, the porous emitter (PE) consists of a packed
bed of randomly arranged spherical alumina particles with average dia-
meter dp ¼ 19 mm and length LPE ¼ 210 mm, and is supported by two
perforated plates made of heat resisting steel plate. The corresponding
optical thickness of the PE is sPE ¼ 2.78. The purpose of the PE is to pro-
vide an energy feedback mechanism by thermal radiation to the PB for
evaporation and serve as a combustor wherein combustion takes place
(partially or wholly) within it.
The combustion characteristics are determined from profiles of the
temperature along the burner axis and the composition of the combus-
tion gases at the exit of the LFPB. In order to know the temperature pro-
files, thermocouples were used with 19 locations of N-type thermocouples
(0.25-mm diameter) inserted through small ports in the burner wall. Each
wire was insulated with magnesium oxide and the entire assembly
sheathed in 1.5-mm diameter. These thermocouples were inserted normal
to the axis of the burner and can be positioned at any radial position
1830 S. Jugjai and C. Pongsai

depending on section (Figure 3). In the PE section, the thermocouples


were positioned at the centerline of the burner, whereas in the PB section
the thermocouples (T1–T4) were positioned at the interface between the
cement lining and the PB so as to prevent the liquid fuel flow pattern
within the PB from being disturbed by protrusion of the thermocouple
junctions. The thermocouple signals are digitized by a general-purpose
data logger, and then transmitted to a personal computer. The thermo-
couple readings can represent the solid phase (or porous medium) tem-
perature since the thermocouples are also solid by themselves. This
simplifies the analysis greatly, and the price we pay for this convenience
is some loss in accuracy due to error arising from several effects, such as
heat transfer effects caused by conduction and radiation etc. To know
the true gas temperatures, the thermocouple readings have to be cor-
rected but this was not performed because radiation error involves
many unknown due to the complex radiative field around the bead
particularly when the thermocouple is inside the porous matrix. More-
over, the uncertainty in the results is high due primarily to the lack of
reliable interface convective heat transfer coefficient and also the lack
of reliable data for thermophysical=chemical and radiative properties
of a porous matrix.
Uncertainties in the intrusive thermocouple measurement are of
concern. However, thermocouple is the only technique available at
present to measure temperature inside a porous medium burner.
Repeated measurement shows an uncertainty of about 10% for the tem-
perature. The results of the measurement are instructive, but should be
interpreted with proper caution regarding the limitations of the experi-
mental technique. Upon steady state condition for each experimental
condition was reached, thermocouple reading for each location was
almost flat with relatively small perturbation. Then the temperatures
were averaged over a specific time interval and plotted on a graph. Error
bars were not employed so as to avoid confusion that may be caused by
several plots being displayed on a single graph. Calibrated rotameters
measured flow rate of the liquid fuel (kerosene) and flow rate of the
combustion air. Combustion pressure DP was measured by standard
manometers.
Emission analysis of the dry combustion products at the LFPB exit is
carried out by using the Messtechnik Eheim model Visit01L, which is a
portable emission analyzer designed especially for quasi-continuous
measurement. A gas processing system of NOx and CO is especially tuned
for electrochemical sensors, ensuring long-time stability and accuracy of
measurement. The measuring range of the analyzer is 0–4,000 ppm
for NOx and 0–10,000 ppm for CO with measuring accuracy of
about 5 ppm (from the measured value) and resolution of 1 ppm for
both NOx and CO. All measured emissions in the experiment are those
Liquid Fuels-Fired Porous Burner 1831

corrected to 0% excess oxygen and dry-basis. Repeated measurement


shows an uncertainty of about 10% for the species concentrations.
The LFPB (SSF and MSF), at a down-flow combustion (h ¼ 0), was
started first by preheating the PB with liquefied petroleum gas (LPG)
before switching to liquid kerosene when the temperature, particularly
at the exit of the PB (T5), attained a favorable value for continuous evap-
oration and auto-ignition (500–700C). The LPG is supplied through the
PB with or without partial mixing with a separate axial air (mixture must
be lean enough to avoid auto-ignition within the PB). Ignition is initiated
by inserting a pilot oxy-acetylene flame through an ignition port at the
side wall of the swirling chamber. Once the ignition is successful, the pilot
flame is withdrawn and the ignition port is closed.
By adjusting flow rate of the axial air and the swirling air with an
appropriate equivalence ratio U, a swirling turbulent flame stabilized
near the exit of the PB (SSF) or matrix stabilized flame (MSF) within
the PE can be achieved. Then, the swirling air is increased to an amount
corresponding to the kerosene flow rate or heat input (CL) to be subse-
quently supplied by admixing with the LPG in the PB. With a suitable
value of U, a kerosene flame can be achieved with the help of heat from
combustion of the LPG=air mixture. Then, the LPG and the axial air
flow rate are gradually decreased until they are completely turned off.
During this course, the overall equivalence ratio U is adjusted by adjust-
ing the swirling air in such a way that stable combustion of the kerosene
flame is maintained while promoting complete combustion. As a steady-
state condition was reached, T, DP and concentration of CO and NOx at
the LFPB exit were recorded before increasing the burner angle h by 15
until h ¼ 180 or an up-flow combustion. Experimental conditions are
summarized in Table 1.

RESULTS AND DISCUSSION

Surface Stabilized Flame (SSF)

Figure 4 shows typical axial temperature profiles of SSF (without PE) at


steady-state conditions. Only 5 temperature profiles, which correspond to
5 burner angles h, i.e., 0, 45, 90, 135 and 180, respectively, are shown
for convenience.
At any burner angle h, turbulent diffusion flames were observed and
occurred almost throughout the cross sectional area of the combustor
under complex flow pattern, i.e., highly turbulent flow created by swirling
flow and impinging flow between jets, resulting in a nearly homogeneous
and complete combustion without soot and odor. The locations of the
1832 S. Jugjai and C. Pongsai

Figure 4. Typical temperature profiles of SSF.

peak temperatures (T7) outside the PB near its exit show locations of the
surface stabilized flames (SSF). In spite of variation in h, the flames were
stabilized and theirs locations were fixed at the same location as the exit
of the swirling air (x ¼ 110). This implies that adequate mixings caused
by highly turbulent flow motion were achieved irrespective of the burner
angle h. Fast evaporation caused by dominating energy feedback mech-
anism by thermal radiation to the PB may also contribute to this flame
stabilization.
Focuses have been made on special features obtained from the present
combustion system; complete evaporation within the PB with temperature
between T4 and T5 being much higher than boiling temperature
(Tboil ¼ 250C) and Leidenfrost temperature (Bernardin and Mudarwar,
1997), efficient preheating effect of the fuel vapor enabling auto-ignition
and temperature profile behaving like a conventional gaseous premixed
free flame with steep temperature gradient across the PB exit.
With steep temperature gradients, heat transfer from flame and from
combustion chamber wall to PB by thermal radiation is responsible for
evaporation and preheating of the fuel vapor within the PB. These
require latent heat and sensible heat, which are supported by interactions
between phases such as convective heat transfer between solid=liquid and
solid=vapor and by thermal radiation absorption by the liquid fuel and
the fuel vapor. This results in a favorably fast evaporation followed by
a nearly homogeneous and complete combustion with almost constant
emissions of CO and NOx of about 150 ppm and 115 ppm, respectively,
as shown in Figure 5.
Liquid Fuels-Fired Porous Burner 1833

Figure 5. Typical CO and NOx of SSF.

Matrix Stabilized Flame (MSF)

With PE installed downstream of PB, it is clear that the combustion


regime has been changed from a surface stabilized flame (SSF) to a
matrix stabilized one (MSF) because the positions of the combustion
zone, where the temperature maxima at any burner angle h are located,
are displaced into the PE (at x  170) as shown in Figure 6. The flames
were stabilized by flame propagation mechanism within PE. Despite a

Figure 6. Typical temperature profiles of MSF.


1834 S. Jugjai and C. Pongsai

smaller heat input CL ¼ 4 kW, MSF yields slightly higher maximum


temperature than that of SSF operating at higher CL ¼ 9.15 kW as
shown in Figure 4. Results of the present study also show that the
maximum temperatures of MSF are higher with high heat input (CL).
This is in agreement with the results obtained by Brenner et al. (2000)
noting the peculiar characteristic of high power-dependent temperatures
of MSF. The corresponding emissions of CO and NOx are shown in
Figure 7. Again, the temperature profiles and the emissions of CO and
NOx were not strongly affected by h the same trend as that of SSF as
shown in Figure 5. However, it seems that MSF system is less affected
by h than SSF because the variation in CO and NOx with h as shown
in Figure 7 is flatter than that shown in Figure 5 for SSF system.
Figure 8 shows typical comparison in temperature profile between
MSF and SSF at exactly the same experimental conditions. Apparently,
MSF yields higher maximum combustion temperature than that of SSF,
confirming the high power-dependent temperatures of MSF (Brenner
et al., 2000). This result can be expected in view of a heat re-circulating
combustion characteristic caused by the additional energy feedback pro-
vided by the thermal radiation emitted from the PE to the PB. In com-
parison with SSF, which is characterized by near adiabatic conditions
and high power-independent temperatures, MSF is strongly influenced
by extensive inter-phase heat exchange within the porous matrix and
by the heat input (CL), resulting in a significant increase in combustion
temperature beyond the normal adiabatic value. The extensive inter-
phase heat exchange leads to an efficient preheating of the unburned
gases, which is automatically organized within a porous medium burner.
Here heat is internally re-circulated to the unburned gases through heat

Figure 7. Typical CO and NOx of MSF.


Liquid Fuels-Fired Porous Burner 1835

Figure 8. Comparison of T.

conduction, thermal radiation characteristics of the solid phase and


redistribution owing to heat exchange between phases. As a result, the
temperature of the combustion of MSF can be significantly improved
as compared with SSF. MSF is, therefore, strongly influenced by heat
input CL. However, experiment of MSF at CL > 6.65 kW at constant
U ¼ 0.52 was not performed because of excessive high temperature and
limitation of the thermocouples.
In spite of higher maximum combustion temperature, MSF yields
lower temperature in the post-combustion zone at the downstream region
of the PE as compared with SSF as shown in Figure 8. This may be
attributed to quenching effect of the hot zone caused by thermal radi-
ation from the PE towards the downstream boundary. This, in turn, leads
to lower average temperature (Tav) along the entire porous burner length,
together with narrowing in the hot zone as compared with SSF. As a
result, MSF yielded relatively lower NOx-emission at almost the same
CO emission as compared with the conventional SSF as shown in
Figure 9, which shows variation in NOx and CO with U. MSF yields
lower NOx than that of SSF throughout the considered range of U.
Again, experiment of MSF at U > 0.52 at constant CL ¼ 6.65 kW was
not performed because of excessive high temperature and limitation of
the thermocouples. MSF can provide the combined effect of lowering
in the average temperature and a shortening in the residence time within
the hot zone, resulting in an efficient measure for suppression of the for-
mation of NOx. The decrease in NOx level by MSF in the present study is
in agreement with other works for similar porous burners (Trimis and
1836 S. Jugjai and C. Pongsai

Figure 9. Comparison of CO and NOX.

Durst, 1996; Brenner et al., 2000). They reported that porous medium
burner can be characterized by low NOx emission because of quenching
effect of the hot zone provided by high heat transport property of the
porous material. Formation of NOx depends on the maximum tempera-
ture, residence time in the hot regions, shapes and transport properties of
the porous medium. In the case of the present study with MSF, quench-
ing effect on the combustion zone caused by presence of the porous
medium could be clearly observed as explained.
Suppression of the formation of NOx by MSF system may also be
attributed to a homogeneous combustion within the PE instead of a het-
erogeneous one as occurs in SSF or a conventional spray combustion sys-
tem. With MSF system, the evaporation zone (PB) is completely
separated from the combustion zone (PE), resulting in a homogeneous
combustible mixture of the fuel vapor and the combustion air to be
formed prior to combustion within the PE. The more homogeneous the
mixture is, the lower the emissions of CO and NOx can be achieved
(Lee et al., 2003) depending on pre-evaporation temperature, mixing pro-
cess (premixed or non-premixed) and supply of the combustion air
(staged or non-staged air). MSF system can provide a more uniformly
distributed equivalence ratio U of the mixture as compared with that of
SSF or a conventional spray combustion system. Thus, a near stoichio-
metric combustion with local equivalence ratio close to one and the
resulting peak temperature can be avoided, even though the combustion
systems are operated at the same overall equivalence ratio.
Figure 10 shows comparison in DP between MSF and SSF. Appar-
ently, DP of the MSF system markedly increases owing to a static press-
ure regain caused by collision of the swirling air on the PE. This increase
Liquid Fuels-Fired Porous Burner 1837

Figure 10. Comparison of DP.

in DP is not a problem so long as it does not have an adverse effect on the


combustion characteristics.
Figure 11 shows a comparison in stable combustion regions between
SSF and MSF at a typical burner angle h ¼ 180. The stable combustion
region was defined as the state where [CO]=[CO2] is lower than 0.002.
[CO] and [CO2], respectively, are mole concentrations of carbon monoxide
and carbon dioxide in dry flue gas at O2 0% base.

Figure 11. Comparison of stable combustion region.


1838 S. Jugjai and C. Pongsai

Undoubtedly, the MSF yields a wider stable combustion region


because it is characterized as a heat re-circulating combustion.

CONCLUSIONS

The major conclusions that can be drawn from the present study are as
follows:
(1) A new combustion system for liquid fuels could be realized by the
proposed LFPB without atomization, offering energy saving for oil
breakup and a high potential to replace conventional spray com-
bustion at any orientation of the burner operating conditions.
(2) By operating the LFPB with the SSF system, relatively low emis-
sions of CO and NOx could be achieved, and these would be further
reduced if operating with the MSF system. The MSF system is
superior to the SSF because MSF can yield a relatively high com-
bustion temperature but with relatively low average temperature,
leading to a favorably complete combustion, and a wide, stable
combustion region. The LFPB with the MSF system could be
classified as a low-NOx burner because its NOx emission is inde-
pendent of the maximum combustion temperature.
(3) Practical applications of the LFPB could include a liquid porous
radiant burner for home heating, a porous combustor-heater for
generating hot water or steam in home use or in industrial applica-
tions, a liquid fuel-fired combustor for a micro gas turbine for elec-
tricity generation, and a hydrogen reformer for fuel cells. However,
much work remains to be done as far as reliability of the products is
concerned. In particular, scaling up in capacity of the burner is very
important for industrial applications. Practical thermal loads could
be met through this scaling-up in capacity of the burner. Also,
theoretical study of this type of burner is a necessary challenge
for thoroughly understanding the phenomena that take place within
the LFPB. This may lead to an accurate scaling-up of the LFPB and
a more efficient utilization of energy.

NOMENCLATURE

CL Heat input, kW
D Diameter, mm
d Inter-distance between PB and PE, or diameter, mm
L Length, mm
MSF Matrix stabilized flame
PB Porous burner
PE Porous emitter
Liquid Fuels-Fired Porous Burner 1839

DP Static pressure difference, mm H2O


SSF Surface stabilized flame
T Temperature, C
x Distance, mm

Greek Symbols
U Equivalence ratio (ratio of theoretical air to practical air supplied)
h Burner angle, degrees
s Optical thickness (a measure of the ability of a medium to attenuate
energy, which is equal to the multiplication between an extinction
coefficient (which is taken here as an absorption coefficient j) and the
geometric length of the PB=PE)

Subscripts
ad Adiabatic
av Average
boil Boiling
PB Porous burner (PB)
PE Porous emitter (PE)

REFERENCES

Bartsch, Ch. (2001) On the way to homogeneous combustion. Motortechnische


Zeitschrift (MTZ worldwide), 62(5), 12–14.
Bernardin, J.D. and Mudarwar, I. (1997) Film boiling heat transfer of droplet
streams and sprays. Int. J. Heat Mass Trans., 40, 2579–2593.
Brenner, G., Trimis, D., Wawrzinek, K., Weber, T., Pickenacker, K., and
Pickenacker, D. (2000) Numerical and experimental investigation of matrix-
stabilized methane=air combustion in porous inert media. Combust. Flame,
123, 201–213.
Durst, F. and Weclas, M. (2001) A new concept of I.C. engine with homogeneous
combustion in a porous medium, Proceedings of the Fifth International Sym-
posium on Diagnostics and Modeling of Combustion in Internal Combustion
Engines (COMODIA 2001), July 1–4, Nagoya, Japan, 467–472.
Fuse, T., Araki, Y., Kobayashi, N., and Hasatani, M. (2003) Combustion char-
acteristics in oil-vaporizing sustained by radiant heat reflux enhanced with
higher porous ceramics. Fuel, 82, 1411–1417.
Fuse, T., Kobayashi, N., and Hasatani, M. (2005) Combustion characteristics of
ethanol in a porous ceramic burner and ignition improved by enhancement
of liquid-fuel intrusion in the pore with ultrasonic irradiation. Exp. Therm.
Fluid Sci., 29(4), 467–476.
Hein, Klaus R.G. (2004) Future energy supply in Europe – challenge and chances.
Fuel, 84(10), 1189–1194.
Jugjai, S. and Polmart, N. (2003) Enhancement of evaporation and com-
bustion of liquid fuels through porous media. Exp. Therm. Fluid Sci.,
27(8), 901–909.
1840 S. Jugjai and C. Pongsai

Jugjai, S., Wongpanit, N., Laoketkan, T., and Nokkaew, S. (2002) Experimental
study on combustion of liquid fuels by a porous medium. Exp. Therm. Fluid
Sci., 26(1), 15–23.
Kaplan, M. and Hall, M.J. (1995) The combustion of liquid fuels within a porous
media radiant burner. Exp. Therm. Fluid Sci., 11, 13–20.
Kaviany, M. (1995) Principle of Heat Transfer in Porous Media, 2nd ed.,
Springer-Verlag, New York, p. 626.
Kayal, T.K. and Chakravarty, M. (2005) Combustion of liquid fuel inside inert
porous media: An analytical approach. Int. J. Heat Mass Transfer, 48(2),
331–339.
Lee, J.C.Y., Malte, P.C., and Benjamin, M.A. (2003) Low NOX combustion for
liquid fuels: Atmospheric pressure experiments using a staged prevaporizer-
premixer. J. Eng. Gas Turbines Power, 125, 861–871.
Liedtke, O. and Schulz, A. (2003) Development of a new lean burning combustor
with fuel film evaporation for a micro gas turbine. Exp. Therm. Fluid Sci.,
27(4), 363–369.
Martynenko, V.V., Echigo, R., and Yoshida, H. (1998) Mathematical model of
self-sustaining combustion in inert porous medium with phase change under
complex heat transfer. Int. J. Heat Mass Trans., 41(1), 117–126.
Park, C.-W. and Kaviany, M. (2002) Evaporation-combustion affected by in-
cylinder reciprocating porous regenerator. ASME J. Heat Trans., 124,
184–194.
Takami, H., Suzuki, T., Itaya, Y., and Hasatani, M. (1998) Performance of
flammability of kerosene and NOx emission in the porous burner. Fuel,
77(3), 165–171.
Trimis, D. and Durst, F. (1996) Combustion in a porous medium-advances and
applications. Combust. Sci. and Technol., 121, 153–168.
Tseng, C.-J. and Howell, J.R. (1996) Combustion of liquid fuels in a porous radi-
ant burner. Combust. Sci. Tech., 112, 141–161.
Weclas, M. (2001) New strategies for homogeneous combustion in IC engines
based on the porous medium technology. ILASS Europe, June 2001.
Weinberg, F.J. (1971) Combustion temperatures: The future? Nature, 233,
239–241.

You might also like