Design of Scrubber PDF
Design of Scrubber PDF
Design of Scrubber PDF
Abstract
Many fuels (oil, wood or process gases) contain components which can be found as acid-forming compounds in the flue gas
after combustion and which are absorbed to a small degree in the condensate when the dew point is reached. This acid, and
therefore highly corrosive condensate, increases the demands on the materials used for the areas affected by condensation in
condensing boilers. The concepts described in the article Design of Scrubbers for Condensing Boilers are based upon use of
the condensate for washing the flue gas. To achieve this, the flue gas is cooled below the dew point in contact with the already
neutralized condensate. This process step allows the wet separation of noxious matter and avoids acid corrosion of the materials
in the area of condensation, raising the choice of possible materials decisively. The review article also exemplifies the state-ofthe-art for condensing boiler technology, as it gives a view of the fundamental principles of two-phase flows, of absorption of
acid-forming gases and their neutralization. Using the examples of sulfuric oxides SOx and nitrous oxides NOx, the effects of
different characteristic features of the gases on the reaction steps are described and possible process steps of the wet separation
in the condensate are discussed. To achieve as high a separation degree as possible between the flue gas and the condensate,
good heat and mass transfer conditions must be guaranteed between the two phasesflue gas and condensate. Dispersing one
of the two phases leads to a strong increase of the interphase. Generally, fluid vaporizers (flue gas as coherent phase) and gas
bubble washers (condensate as coherent phase) can be taken into consideration. Advantages and disadvantages of these
absorbers are worked out for use as washers in combination with condensing boiler technology and the fluid-mechanical
principles necessary for the design of a gas bubble washer. The sometimes contrary influences of constructive parameters
on pressure changes in the flue gas, as pressure loss and mass transfer conditions between flue gas and condensate, are
discussed. q 1999 Elsevier Science Ltd. All rights reserved.
Keywords: Condensing boiler; Scrubber; Desulfurization; NOx absorption; Neutralization
Contents
Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2. State of science and technology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1. Condensing boiler technology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.1. Dew point and combustion efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.2. Integration of condensing boiler technology in the heating system . . . . . . . . . . . . . . . .
2.1.3. Concepts for condensing boiler technology for sulfurous fuels . . . . . . . . . . . . . . . . . . .
2.2. Deposition of sulfur oxides from flue gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.1. Phase equilibria and oxidation at the absorption of SO2 . . . . . . . . . . . . . . . . . . . . . . . .
2.2.2. Mass transfer at SO2 absorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
306
308
308
308
308
311
312
314
315
316
* Corresponding author. Address for correspondence: Shell Research and Technology Centre, Hamburg, Deutsche Shell AG, PAE Labor,
OGMPT/4, Hoh-Schaar-Str. 36, 21107 Hamburg, Germany. Tel.: 1 49-40-7565-4739; fax: 1 49-40-7565-4581.
E-mail address: frank.f.haase@ope.shell.com (F. Haase)
0360-1285/99/$ - see front matter q 1999 Elsevier Science Ltd. All rights reserved.
PII: S0360-128 5(99)00002-7
306
2.2.3. Neutralization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3. Wet separation of nitrogen oxides from flue gases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.1. Solution of NO and NO2 in water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.2. Oxidation of NO in aqueous solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.3. Absorption and reduction of NO in aqueous solution . . . . . . . . . . . . . . . . . . . . . . . . . .
3. Experimental apparatus and measuring techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1. Design of the test boiler . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2. Measuring techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.1. Gas analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.2. Water analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4. Fluid-mechanical layout of a gas bubble washer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1. Gas flow through the complete perforated bottom . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2. Drag reduction of the flue gas flowing through the water bath . . . . . . . . . . . . . . . . . . . . . . . . .
4.3. Mass transfer between dispersed gas and liquid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5. Experiments on the absorption and neutralization of SO2 and NOx in the water bath . . . . . . . . . . . . .
5.1. Neutralization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.1. Formation of Mg(OH)2 via hydration of MgO . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.2. Carbonate formation by CO2 absorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2. SO2 absorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2.1. Influence of pH on the SO2 absorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2.2. Sulfite oxidation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.3. Wet absorption of nitrogen oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.3.1. Oxidation and absorption of NO in the water bath . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.3.2. Absorption of NO2 in water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Nomenclature
Physical symbols
A
area (m 2)
a
volume specific interphase (m 21)
b
weir height (m)
c
concentration (mg/l, mol/m 3)
c
specific heat capacity (J/kg K)
d
diameter (m)
dS
Sauter diameter (m)
E
specific emission (mg/MJ)
F
force (N)
f
power ratio of a radiator
Fr*
modified Froude number
g
gravitation constant (m/s 2)
H
enthalpy (J)
h
height of water column above perforated
bottom (m)
H
Henry constant (N m/mol)
h
specific enthalpy (J/kg)
hfg
latent heat of vaporization (J/kg)
HHV
higher heating value (J/kg)
K
equilibrium constant (mg/l) 21
k
velocity constant (s 21)
l
length (m)
LHV
lower heating value (J/kg)
m
mass (kg)
m_
M
n_ 00
318
319
319
320
321
322
322
323
323
323
323
324
327
327
329
329
330
330
332
332
332
334
334
335
335
336
v_d
V_
Symbols
w
We
N
P
p
pH
Q
q_ 00
Q_
Q_ std
R
r
r
s
T
t
t90
x
y
z
a
b
D
1i
1
h
l
q
Dq ln
r
s
j
c
mole fraction
height of perforated bottom above boiler
bottom (m)
number of holes
absorption degree
mass transfer coefficient (mol/N s, m/s)
difference
Henry activity coefficient
volume fraction
efficiency degree
airfuel ratio
temperature (8C)
logarithmic average temperature (K)
density (kg/m 3)
interfacial tension (N/m)
mass fraction
mole fraction
Chemical symbols
Ag
silver
AgCl
silver chloride
Al
aluminum
Ca
calcium
Ca(OH)2
calcium hydroxide
CaCO3
calcium carbonate
CaO
calcium oxide
CaSO3
calcium sulfite
CaSO4
calcium sulfate
methane
CH4
ClO42
perchlorate ion
CO2
carbon dioxide
carbonate ion
CO322
e
electron
EDTA
ethylene diamine tetra-acetate
Fe
iron
H1
hydrogen ion
H2CO3
carbonic acid
H2O
water
H2O2
hydrogen peroxide
H2SO3
sulfurous acid
H2SO4
sulfuric acid
HN(SO3)222 imido disulfonate
HNO2
nitrous acid
HNO3
nitric acid
HO22
hydrogen dioxide ion
HON(SO3)222 hydroxylamine disulfonate
H3PO4
phosphoric acid
hydrogen sulfite ion
HSO32
K
potassium
KClO4
potassium perchlorate
potassium permanganate
KMnO4
M1
metal cation
M2SO3
metal sulfite
Mg
magnesium
Mg(OH)2
magnesium hydroxide
MgNO3
magnesium nitrate
MgO
MgO2
MgSO3
MgSO4
Mn
MnO2
MOH
N(SO3)332
N2
N2O
N2O4
Na2O2
Na2C2O6
Na2SO3
Na2SO4
NaOH
NH3
(NH4)2SO4
NO
NO2
NO22
NO32
NOx
NTA
O2
O3
OH 2
PbO2
Pt
Rh
S2O422
S2O622
SiO2
SO2
SO3
SO322
SO422
magnesium oxide
magnesium peroxide
magnesium sulfite
magnesium sulfate
manganese
manganese dioxide
metal hydroxide
nitrilo trisulfonate
nitrogen
dinitrogen monoxide
dinitrogen tetroxide
sodium peroxide
sodium peroxocarbonate
sodium sulfite
sodium sulfate
sodium hydroxide
ammonia
ammonia sulfate
nitrogen monoxide
nitrogen dioxide
nitrite ion
nitrate ion
nitrogen oxide
nitrilotriacetate
oxygen (molecular)
ozone
hydroxide ion
lead dioxide
platinum
rhodium
dithionite ion
dithionate ion
silicon dioxide
sulfur dioxide
sulfur trioxide
sulfite ion
sulfate ion
Index
*
0
aq
b
bu
c
c
crit
d
dp
fg
fp
g
g
gran
h
i
state of saturation
standard conditions
dissolved
base
gas bubble
combustion
continuously
critical
dispersed
dew point
flue gas
flow pipe
gas side
gaseous
granulated
heating medium
component
307
308
in
kin
l
l
lim
max
min
out
p
r
s
std
wb
entering
kinetic
lifting force
liquid side
limit
maximum value
minimum value
leaving
pressure
return
interfacial tension
standard state
water bath
1. Introduction
Burning fossil energy sources, the acid-forming components sulfur dioxide (SO2) and nitrogen oxides (NOx)
are formed, as well as others, and are emitted into the atmosphere. Following the mass flow of the nitrogenous and
sulfurous compounds after their emission, it is necessary to
differentiate between their effects on the air and on plants,
earth and water.
As the average retention times of the trace gases SO2 and
NOx are only very short (hours up to several days), their
influence on air quality is regionally limited. SO2 and NOx
are toxic to human beings at a certain concentration. The
current discussion about anthropogene influences on the
climate considers SO2 and NOx not in the first place. The
major climate influencing gases in the atmosphere are steam
(H2O), carbon dioxide (CO2), ozone (O3), dinitrogen oxide
(N2O) and methane (CH4). These gases cause the so-called
greenhouse effect of the atmosphere. But SO2, NOx (and
ammonia NH3) lead to the formation of aerosols which are
solid or liquid parts in the air outside of clouds with a
diameter between 0.001 and 100 mm. They scatter and
absorb solar irradiation and emit thermal radiation so they
behave contrary to the anthropogene heating of the earth [1].
The largest part of the emission is chemically converted
in the air before it is deposited via rain. At the chemical
conversion of sulfur dioxide, sulfuric acid (H2SO4) is
formed; its salts are called sulfates (SO422). Finally, gypsum
or Na2SO4 is formed. Nitrogen monoxide is oxidized to
nitrogen dioxide, which forms nitric acid; its salts are called
nitrates (NO32).
Before the acids formed by SO2 and NOx get into the
ground, the protective wax coating on tree leaves and
needles can be damaged. Nutritive substances are washed
out. Also, the supply of nutritive substances of the ground
decreases. With increasing acidity of the ground, first
calcium magnesium and potassium, and in later stages,
manganese and aluminum are dissolved out of the ground
particles. They are washed out and lost to the ecosystem [2].
Together with the washed out alkali sulfates and nitrates
309
Fig. 1. Deposition of SO2 and NOx in the condensate of the flue gas as an alternative to their emission.
310
Fig. 2. Dew point temperature q dp of the flue gas from the combustion of heating oil EL versus airfuel ratio l .
Fig. 3. Increase of the dew point caused by sulfuric acid in the flue gas (calculated data) [3].
311
flue gas. Good mass transfer conditions between flue gas and
necessary condensate result in complete saturation of the
flue gas with steam, so that for only a slight increase of
the temperature above the dew point, condensate will vaporize and combustion efficiency decreases compared to a dry
flue gas system. This phenomenon is shown in dashed
curves in Fig. 4 and provides the necessity of avoiding
flue gas temperature above the dew point by appropriate
control systems.
2.1.2. Integration of condensing boiler technology in the
heating system
For the layout of a heating system, the so-called standard
heat demand of a building, Q_ std is used; it describes the heat
power supplied to heat the rooms to comfortable temperatures at a low outdoor temperature [4]. Using the standard
heat demand for the layout of the boiler (provided that the
piping losses are negligible), the boiler has to supply the
power
3
Q_ h Q_ std m_ h ch qfp 2 qr
to the heating circuit. In the aforementioned equation, mh
represents mass flux of the heating medium, ch the specific
heat capacity of the heating medium, q fp the flow-pipe
temperature of the heating medium and q r the return
temperature of the heating medium.
The values of the process parameters mh, q fp, q r and their
combinations can only be chosen within certain limits. If the
flue gas is cooled below its steam dew point by the heating
medium, the maximum return temperature is then determined according to Fig. 2 depending on the airfuel ratio
l chosen.
Fig. 5. Influence of flow-pipe and return temperature on the power ratio f of a radiator.
312
Q_ , A Dqln n
4
qfp 2 qr
ln qfp 2 qroom =qr 2 qroom
The value of the exponent n lies within the limits 1.1 to 1.5;
for radiators, n < 1.3 is obtainted [5], for floor radiators, n <
1.1 [6]. Making the heat flux from Eq. (4) specific to standard conditions (q fp,std 908C, q r,std 708C, q room,std
208C) [5], the so called power ratio of a radiator for equalsized radiator surfaces
!n
Q_
q_
Dqln
Dqln
f _
6
q_std
Dqln;std
59:44 K
Qstd
This relation is valid for m_ H =m_ H;N $ 0:2 and can be used to
determine the heat a radiator supplies for various pairs of
flow-pipe and return temperature. Fig. 5 shows the power
ratio f dependent on the return temperature q r for different
flow-pipe temperatures q VL. It can be seen from Fig. 5 that
the supplied heat flux of a radiator decreases
if the temperature difference (q fp 2 q r) is increased
when the return temperature is lowered by reducing the
heat medium mass flux and keeping the flow-pipe
temperature constant (line A),
if the temperature difference is decreased by lowering
the flow-pipe temperature, keeping the return temperature constant (line B),
if the temperature level is decreased, keeping the
temperature difference constant (line C).
If the same heat flux is to be obtained as in the standard
case, the radiator surface must be increased according to
A
1
A :
f std
313
Fig. 9. Condensing boiler technology with flue gas desulfurization; concept IV [16].
314
For gas temperatures higher than 12008C, no relevant oxidation takes place because of the state of the equilibrium.
Also, no relevant oxidation can be observed when cooling
the flue gas in the boiler, because the velocity of reaction (8)
is low [18]. SO3 shows a very good solubility in water and
reacts under the presence of steam to form sulfuric acid.
Thus, the technology used for a deposition of SO2 generally
also achieves good results for SO3, therefore in the following, the deposition of SO2 will be seen as the central point of
the investigations. In the USA and in Japan, flue gas desulfurization units have been used in power plants for over 30
years; in Germany, their use began about 20 years ago [19].
The long development time has led to a number of process
variants which generally can be divided into washing, spray
absorption and drying processes.
Washing processes
Washing processes usually work with suspensions of
calcium oxide (CaO, caustic lime), calcium hydroxide
(Ca(OH)2, slaked lime) or calcium carbonate (CaCO3, limestone). The primary product of the lime washing process is
calcium sulfite (CaSO3). In Germany, gypsum CaSO42H2O
is usually gained via oxidation, which is either dumped or
made usable by subsequent processes [20]. Alternatively,
the Walther Company uses an ammonia solution for washing and acquires ammonium sulfate ((NH4)2SO4) as a
product which can be used as synthetic fertilizer [21]. As
a result of the good solubility of the ammonia salts, no
fouling problems occur. MontanWerke Brixlegg, in
Austria, uses a washer which is operated with an
Mg(OH)2 suspension which is produced from MgO by
adding water [22].
Spray absorption processes
During spray absorption processes (also called mixed methods), calcium hydroxide suspensions with a high solid fraction are sprayed into the hot flue gas stream [23]. The water
vaporizes and the salts formed are drawn off, together with
the excess lime and some fine dust, by textile or electric
filters positioned further downstream. A disadvantage is
the over-stoichiometric consumption of lime and the lower
deposition degree of SO2 as compared with wet processes.
Dry processes
During dry processes, we differentiate between lime absorption (dry additive process) and the absorption via activated
coke (mining technology). During dry additive processes,
small-grained lime is blown into the hot flue gas. As an
alternative, lime may be added to the fuel. The calcium
sulfite, calcium sulfate and the remaining lime which are
formed are drawn off a dust filter. Low quality coal and
brown coal naturally contain lime. In pulverized coal
furnaces, which are common in boilers of large dimensions,
lime reaches the combustion chamber as fine dust and
absorbs part of SO2 in the ash. A complete desulfurization
seems possible with the fuel-own lime if a low combustion
315
Fig. 10. Course of concentration and partial pressure of SO2 according to the two-film theory [27].
10
11
316
12
13
15
16
1
O aq N SO22
4
2 2
17
reduction
N oxidizer 1 electron
oxidation
18
1
O 1 2e2 1 2H1 N H2 O
2 2
SO22
3 1
1
O 1 2OH2 1 2H1 N SO22
4 1 2H2 O
2 2
19
20
21
22
23
24
n_ 00SO2
where
is the mole number per area and time of the
transferred SO2 b g, b l mass transfer coefficient for the gas
side in mol/Ns, for the liquid side in m/s.
Further
n_ 00SO2
dn 00SO2
dpSO2 1 Nbu
2
dt
dt pg Abu
25
where Nbu is the mole number in the gas bubble, Abu the
surface of the gas bubble and pgtotal pressure of the gas in
the bubble.
Substituting Eq. (22) in Eq. (25) and integrating over time
within the limits of entering the liquid and leaving the
liquid, we obtain:
pSO2 2 pSO2 ;int uttout
26
A
pSO2 2 pSO2 ;int uttin exp 2bg pg bu tout 2 tin
Nbu
27
317
28
30
318
Table 1
Solubility of the compounds at 20 and 508C in 100 g water
Cation
Reactants
Hydroxide (g)
Products
Carbonate (g)
Sulfite (g)
Sulfate (g)
Na
208C
508C
Mg 21
208C
508C
Ca 21
208C
508C
a
b
52.2 [30]
59.2 [30]
18.1 [30]
32.2 [30]
27.1 [30]
34.7 [30]
0.000017 a [30]
0.000015 b [30]
0.000072 a [31]
0.126 [18]
0.0917 [30]
0.0065 [30]
0.0038 [30]
0.0043 a [30]
0.0057 [30]
At 188C.
At 458C.
Eq. (29) makes clear that even under ideal mass transfer
conditions, the SO2 partial pressure after the washer is not
zero; so the conversion of the sulfite in the liquid is of a
particular importance.
2.2.3. Neutralization
By absorption of sulfur dioxide or other acid gases, the
pH decreases and the adsorbent will be quickly saturated
with SO2. Raising the pH to a neutral area (neutralization)
will change the equilibria described in Section 2.2.2, so SO2
can be further dissolved. The neutralizer must either be
added continuously during the absorption of SO2, or it
must be in excess in the absorbent.
During neutralization, an acid reacts with a base under the
formation of a salt and water. The main reaction is the
synthesis of hydrogen ions and hydroxide ions:
H1 1 OH2 N H2 O
31
32
33
34
319
35
is exothermic, NO does not react with oxygen at temperatures above 6508C [18]. Nitrogen oxides emitted from
furnaces primarily exist as NO.
2.3.1. Solution of NO and NO2 in water
Gas is well absorbed in liquid, especially if the gas and
the liquid show about the same polarity. The polarity of a
molecule is determined by its structure and the difference
between the electronegativity of the atoms in the molecule.
As a result of its angular structure, water shows a large
dipole moment and therefore can easily dissolve molecules
with a high polarity.
Nitrogen monoxide (NO) has a low solubility in water as
a result of its low polarity. The linear structure of the molecule and the relatively small differences between the electronegativity of oxygen and nitrogen result in NO, which
320
36
37
1
2
1
HNO3 1 NO 1 H2 O
3
3
3
38
39
40
41
42
43
321
Fig. 13. Basic layout of the test boiler with water bath [37].
44
45
2
SO2 1 SO22
3 1 H2 O ! 2HSO3
46
22
2FeIIEDTANO 1 2SO22
3 ! 2FeIIEDTA
1 N2 1 2SO22
4
47
322
Table 2
Hole diameter d0, number of holes z and minimum volume flow
V_ bottom;std of the perforated bottoms used
Perforated Bottom
d0 in mm
1
2
3
4
5
2.5
2.5
1.5
3.5
2.5
z
80
120
192
40
128
V_ bottom;std in m 3/h
12.0
18.0
11.0
11.0
19.2
Table 3
Manufacturers data of the gas analyzers used, NDIR [38], electrochemical measuring cell [39] and chemiluminescence [40]
Measuring value
Measuring method
Measuring range
Linearity
Resolution
Exactness
T90 time
cSO2
NDIR
01000 ppm
^20 ppm
2 ppm
5 ppm
160 s
Electro-chemical cell
020 ppm
no data
0.2 ppm
2% of measured value/month
60 s
cNO ; cNOx
Chemiluminescence
0100 ppm
^1 ppm
0.1 ppm
1 ppm
1s
cCO2
NDIR
020%
^0.4%
0.05%
0.1%
160 s
cCO
NDIR
0300 ppm
^6 ppm
2 ppm
1.5 ppm
160 s
323
324
other for the gas volume flow V_ c . They are separated only by
a small liquid lamella.
Until this volume flow is reached, single bubbles are
formed. If the volume flow is increased further, the thin
liquid lamella is destroyed and a jet is formed which leads
to the formation of bubble clusters with bubbles of different
sizes.
The limit between the formation of single bubbles and
bubble clusters can be determined by comparing the forces
at the liquid lamella [44]. It is assumed that the liquid shows
only a low viscosity. From the side of the hole, the kinetic
force Fkin and the force of interfacial tension Fs must be
considered. From the side of the gas jet, we take the pressure
force Fp, arising from the curved surface, and the lifting
force Fl on the liquid lamella into account. These forces
can be determined as follows:
p
Fkin rd w20 d02
48
4
Fs psd0
49
Fp
p
sd
2 0
50
Fl
p
r 2 rd gsd02
4 c
51
325
Fig. 15. Calculated Weber numbers We, Wecrit and measured critical Weber number Wecrit for a just completely passed perforated bottom, d0
2.5 mm [37].
where
We
53
p
p
p
) psd0 1 rc 2 rd gsd02 rd w20 d02 1 sd0
4
4
2
2s
r 2 rd
1 c
rd
rd w20 d0
2
1 4=5
1
0:45395 1
We
Fr*
d0 g
w20
0:45395 1
57
5=4
w20
rd
:
d0 g rc 2 rd
58
and
54
Fr*
!4=5
rd w20 d0
s
55
56
*
*
Fig. 16. Calculated modified Froude numbers Fr*, Frcrit
and measured critical modified Froude number Frcrit
for a completely passed perforated
bottom, d0 3.5 mm [37].
326
Fig. 17. Drag reduction of the flue gas flowing through the water bath [37] perforated bottom 1: z 80, d0 2.5 mm, V_ bottom;std 12 m 3/h
perforated; bottom 2: z 120, d0 2.5 mm,V_ bottom;std 18 m 3/h.
Fig. 18. Drag reduction of the flue gas by bubble formation and by the perforated bottom perforated bottom 3: z 192, d0 1.5 mm, V_ min
11 m 3/h, perforated bottom 4: z 40, d0 3.5 mm, V_ min 11 m 3/h.
At the layout of a perforated bottom for flue gas dispersion, amount and diameter of the holes must be chosen so
that the bottom is passed completely by the gas. If too many
holes (or too big in diameter) are chosen, the bottom is not
passed completely by gas and inactive zones will result,
while for too few holes, an increase of the gas volume
flow means an additional drag reduction while passing the
hole, which results in the formation of larger bubbles.
Choosing burner power, airfuel ratio and flue gas temperature before entering the water bath determines the volume
flow of the gas.
Figs. 15 and 16 show the experimental results for completely passed perforated bottoms of different geometry. For
these experiments, the test boiler shown in Fig. 13 was used.
To allow good observation of the bubble formation, the heat
exchanger was not built, and a plexiglass boiler wall was
used.
In Fig. 15, the Weber number We for a completely passed
bottom is the determining number, as small holes were
chosen with a diameter d0 2.5 mm. The curves show the
calculated Weber number as a function of the number of
holes and volume flow of the gas V_ bottom;std which flows
through the bottom under the conditions (q std 258C,
pstd 1 bar). The measured points as shown represent the
volume flow of the gas for which the bottom is passed
completely by the gas for the chosen number of holes z.
In Fig. 16, the modified Froude number Fr* is used for the
determination of a perforated bottom, as large holes were
chosen with a diameter d0 3.5 mm.
*
The values of the critical numbers Wecrit and Frcrit
, determined in the experiments, are higher than the theoretically
*
*
calculated values Wecrit
2 and Frcrit
0.37. The difference
between theoretical and experimental values increases with
a growing volume flow V_ bottom;std . This phenomenon can be
explained because the perforated bottom does not cover the
complete cross-section of the liquid. The liquid taken into
motion by the rising gas bubbles is diverted when it reaches
the liquid surface, and it flows back downward in the section
which is not passed by the gas (d2 , d , d1). For such a
distribution of the dispersed gas over a cross-section of the
liquid, a stationary flow of the liquid will result, which we
call circulating flow. As a result of this circulating flow, a
fifth force at the liquid lamella must be taken into account,
which supports the lifting force Fl and therefore is a
constructive force for the lamella. Thus, bubble cluster
*
formation begins at higher gas velocities, so Wecrit and Frcrit
increase. With an increasing gas volume flow V_ bottom;std , the
*
circulating flow becomes stronger and Wecrit and Frcrit
increase.
The formation of the circulating flow is enforced by the
geometrical arrangement of the perforated bottom to
achieve a stable flow of the liquid. If the complete crosssection of the fluid is covered by the perforated bottom,
rising and falling liquid flows come in contact with each
other. Even the smallest disturbances can then lead to a
complete breakdown of the stationary flow and can
327
328
Fig. 19. Influence of MgOgran hydration on the pH [37], m_ Oil 1.0 kg/h, j S 0.12%, mMgO 1582 g.
1d
dS
60
V_ bottom;std
:
A
62
63
329
Fig. 20. Dissociation equilibria of CO2, HCO32 and CO322 versus pH and base fraction xb.
V_ bottom;std
0:04 m=s
64
330
Fig. 21. Absorption of CO2 from the flue gas in the water bath for a stepwise increasing pH by caustic soda.
65
With the absorption of the acid-forming flue gas components, a strong decrease of the pH, down to pH 5.7,
can be observed within the first three minutes of the first
burner operation sequence. Having reached this minimum,
the pH increases slowly until it reaches a value pH 5.6, at
the end of the operation sequence. Within the first 3 min of
the burner operation sequence, the hydroxide formed before
is consumed for the neutralization of the formed acid. The
MgO which was covered with Mg(OH)2 before the start of
the burner is set free and the velocity of the hydroxide
formation increases, so the pH increases as well.
After 72 h, the burner is started again. The pH decreases
slower than at the first operation sequence and reaches its
minimum at pH 6.8 before increasing to its final value
pH 6.9. The reason for the slower decrease of the MgO
and the less significant minimum is to be seen in the higher
amount of Mg(OH)2 that has been formed within the 72 h.
Also, the following operation sequences, 96 and 102 h later,
show a similar course of the pH. The stationary pH at the
end of each operation sequence increases only slightly with
increasing employment time of the MgO in the water bath,
and remains below pH 7.5 even for an operation time of
several weeks.
The pH at the start of each burner operation sequence (t
0) decreases with the total burner operation time. The reason
for this is the increasing concentration of Mg 21 in the water
bath. During the standstill times of the burner, magnesium
hydroxide dissolves until the solubility product of Mg 21 and
OH 2 is reached. Thus, an increasing concentration of Mg 21
has a higher part on the solubility product and the OH 2
331
Fig. 22. Dissociation equilibria of SO2, HSO32 and SO322 versus pH and base fraction xb.
66
Fig. 23. Absorption of SO2 versus pH and height of the water above the perforated bottom h [37], m_ Oil 1.0 kg/h, j S 0.12%.
332
Fig. 24. O2 concentration in the water bath versus temperature q wb and airfuel ratio l [16].
333
Fig. 25. NO absorption degree a NO versus pH and H2O2 concentration at experiment start [37], q wb 358C.
Fig. 26. Specific emissions ENO before and after the water bath in dependence on the airfuel ratio l [37], mMgO2 270 g, pH 7.5, h 10 cm.
334
Fig. 27. Absorption of NO2 versus height of the water column and the NO2 start concentration when entering the water bath; pH 7.
1
O N SO22
4
2 2
69
70
2
HO2
2 1 NO N NO2 1 OH
71
2
2HO2
2 N O2 1 2OH
72
H3 O1 1 OH2 N 2H2 O
73
335
6. Summary
Many energy sources, like oil, wood or process gases,
contain components which show acid-forming compounds
in their flue gas after combustion, and which, to a small
degree, are absorbed in the condensate when the dew
point falls short. The acid so-formed, and highly corrosive
condensate, raises the requirements on the materials used in
the condensation area.
The concepts described in this paper are based on the idea of
using the condensate for washing the flue gas; thus the flue gas
is brought in contact with an already neutralized condensate
and thereby is cooled below its dew point. This step process
allows the wet separation of noxious matter and avoids acid
corrosion of the materials used in the condensation area.
Thereby, the choice of usable materials is increased.
With the example of a gas bubble washer, fluid-mechanical investigations of the flue gas dispersion in the water
bath and experiments on the wet separation of SO2 and
NOx are presented. The absorption of the gases and neutralization of the formed acids was regarded at the wet
separation, whereas we primarily investigated the absorption of SO2.
The fluid-mechanical investigations on the flue gas
dispersion in the water bath were aimed at realizing good
mass transfer conditions between flue gas and water, and a
small expense of energy for the realization of the dispersion.
The dispersion of the flue gas in the water is achieved by a
perforated bottom. Its design, referring to the number of
holes, is determined via the critical Weber number for
small holes and via the critical modified Froude number
for large holes. For the parameters chosen, a minimum gas
volume flow is given.
To enforce a circulating flow of the water, only a part of
the water bath cross-section is covered by the perforated
bottom. Above the perforated bottom, water flows in an
upward direction and is diverted when reaching the water
336
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
337