Você está na página 1de 18

OPTIMISED INTEGRATED FGD PROCESS

A. STERGARŠEK1, M. GERBEC1, R. KOCJANČIČ1, and I. HRASTEL2


1
"Jožef Stefan" Institute, 2ESOTECH d.d.
andrej.stergarsek@ijs.si

ABSTRACT
The work on further development of wet calcite flue gas desulphurization (FGD) with the main focus on
low-cost process is described. The whole process has been critically analyzed, including physical and
chemical parameters of the SO2 capturing and equipment layout.
Two physical-chemical non-stationary models have been developed that can take into account input
operating parameters and give energy consumptions, absorption efficiency and waste gypsum slurry
composition. One model describes spray scrubber and the other the falling film absorber. The
optimization presented shows that falling film absorber is less energy consuming alternative. The results
of scaling formation experimental study are presented.
Other field of optimization is the layout of the equipment and the integration of sub-processes in the
same equipment. Some integration steps proposed, designed and tested are as follows:
• Intermediate product (gypsum) storage is done in the absorber make-up tank, by allowing the
density to rise up to 20 % solids.
• Calcite slurry is instantaneously produced from calcite powder wetted by absorption slurry.
• Hydrocyclons are fed directly from re-circulating pumps and underflow is fed directly to the
automatic filter or centrifuge.
Detailed case study was prepared for a 215 MW PP. Up to 30 % lower investments can be expected in
comparison with current offers on the market. In respect to operating costs, the energy consumption and
cost of maintenance could be decreased by 10 to 25%. As shown by risk analysis the availability of the
system remains equally high as for classical lay-outs.

KEY WORDS: FGD modelling, optimisation, equipment integration

INTRODUCTION
The general objective of the work was to develop non-stationary models of the wet calcite flue gas
desulphurization process and use the models to optimise the process in respect to investment and
operating costs. Cost reductions will be achieved through the development of some critical parts in
chemical process of desulphurisation and through the optimisation of the process and equipment layout.
At the present level of development of FGD process, the most promising approach for absorber
optimisation (being it spray tower design or falling film design utilizing packed internals) is the use of
specialized theoretical models which consider gas phase aerodynamics, mass transfer of SO2 across
gas/liquid phases, and chemical reactions involved (absorbed SO2 dissociation, sulphite oxidation rate,
solid neutralization agent-calcite dissolution, gypsum crystallization, etc.). Models were applied to the
design of case study FGD plant and comparative results are given. Scaling formation in the absorber
internals was investigated in small pilot continuous FGD unit, the results are very important in falling
film applications.

FGD MODELLING
Non-stationary FGD models were developed for spray scrubber and falling film absorber, respectively.
The following sub-models were developed and integrated into model calculations:

1/18
i) Hydrodynamics of the drop and mass transfer from gaseous to liquid phase (drop).
ii) Hydrodynamics of the falling film of absorbing slurry and mass transfer from gaseous to
liquid phase
iii) Chemical equilibrium in liquid phase
iv) Dissolution of solid neutralisation agent, limestone
v) Oxidation of sulphite (experimental model).
vi) Precipitation of produced gypsum.

Spray tower modelling, measurements and optimisation

Hydrodynamics of the drop and mass transfer from gaseous to liquid phase (drop).
According to the investigation of Kronig and Brink (1) the following flow pattern is established inside a
freely falling drop: the surface layer of the liquid moves parallel to the movement of the surrounding
medium, while the liquid at the stagnant point of the gas emerges from the inner part, and on the “lee”
side the surface layer dives back into the drop. Deeper layers of liquid move parallel to the outer layer
(Figure1). In the cross-section of the drop we can observe a stagnant point, at a distance of 0.707 radii
from the drop centre (perpendicular to the direction of non-disturbed flow of the surrounding media).
Investigation on single drop of absorbing liquid phase described elsewhere led us to a simplified model
of the internal circulation and to make it useful for the general non-stationary model. The following
internal flow pattern as shown in Figure 1 was set up: a uniform outer layer of liquid is driven by the gas'
kinetic energy. At the “lee” point, the liquid dives into the kernel of the drop, into which it flows in a
“plug-flow” manner, leaving the kernel at the point of gas stagnation. For the sake of calculation, the
kernel and the outer layer are divided into equivolume increments of liquid. In these increments, all
chemical equilibria are calculated as they proceed in time.

A B
LEE SIDE

EQUIVOLUME
INCREMENTS

0.7 r
r r
POINT OF
STAGNATION OF GAS

RELATIVE MOVEMENT
OF NON-DISTURBED CONTINUUM (vG )

Figure 1: Flow pattern in the drop according to the investigation of Kronik and Bbrink, A, and
simplified model with internal plug flow used in this model, B

2/18
In mass transfer calculations of the amount of SO2 absorbed (n(SO2 iter)) at given model point in the
absorber overall resistance on the interface (KG), interfacial area (Aiter), model iteration time (titer) and
local gas phase concentration difference were considered for each iteration:

*
n(SO2 iter) = K (G iter) t iter Aiter ( P(SO2 iter) - P(SO2 iter) )

*
P(SO2 iter) = H SO 2 c( SO2 L )

with P(SO2 iter) as gas phase concentration and P*(SO2 iter) as corresponding gas phase concentration from
Henry's law and local SO2 concentration in liquid phase (c(SO2 L), calculated from chemical equilibria).
HSO2 is a Henry's constant for SO2/H2O equilibrium and was calculated according to Smolik and Vitovec
(3).
The rate of SO2 transfer across a gas/film interface (overall coefficient KG) was calculated from both gas
(kG) and liquid side (kL) transfer coefficients and from additional enhancement by a chemical reaction
taking place after the absorption of SO2:

1
KG =
1
+ H SO 2
k G β SO 2 k L

βSO2 is an absorption enhancement factor due to chemical reaction, which was calculated according to
Carmichael and Peters (4). It does not consider the differences in diffusivities of dissolved species of
SO2, the effect of SO2 oxidation and reaction with dissolved carbonates:

S4 S4 S4
β SO2 = 1 + K 1 + K 12 K 2
c( H 3O+ ) c( H 3O+ )

Chemical equilibrium in liquid phase


The chemical equilibrium in the liquid phase was modeled as reported by Stergaršek et al. (2) and has
included absorption, hydrolysis, dissociation, desorption, oxidation, neutralisation of all relevant
chemical species. Again, the temperature and the activities of the ions in the solution were considered.
The equilibrium constants at a given temperature were calculated using the published data, Brewer (5),
and the activity coefficients were calculated for more concentrated systems according to the
Pitzer/Debye/Hückel equation as used by Luckas et al. (6):

1
Ix= ∑ j z 2j x j
2

log( γ z )=
Aφ   2 z 2 
  ( ς 1/2
) (z 2 I 1/2x - 2 I 3/2x )

(1 + ς I 1/2x ) 
ln 1 + +
M H1 / 22O   ς 
Ix

where xj is the mole fraction of ion j in solvent, z is charge of ion j, Aφ and ς are Debye-Hueckel
parameters and MH20 is the molar mass of water.

3/18
Dissolution of solid neutralisation agent, limestone
Ground limestone (CaCO3) is used as a neutralizing agent. It partially dissolves in water, releasing Ca2+
and CO32- ions:
CaCO3 (s) ↔ Ca + CO3
2+ 2-

The limestone dissolution rate was measured in a pH-stated agitated tank, at different pH values,
temperatures and CO2 partial pressures in the gas phase. To simulate typical conditions in the FGD
application, a 0.1 mol L-1 CaCl2 concentration in a bulk solution was used. A well-oxidized FGD
conditions were assumed and thus limestone blinding due to sulphite precipitation was not considered.
The experimental dissolved limestone amount time dependence was simulated using a known limestone
sample particle size distribution by finding the spherical particle(s) linear wall dissolution rate (kCaCO3).
For each experiment, the best fit value of kCaCO3 at given conditions was found by fitting model results to
the experimental data.
As a result of our experimental work an empirical correlation was produced to calculate the linear
dissolution rate of the limestone particles depending on pH value, temperature and CO2 concentration in
the gas phase, Gerbec (7,8):

 
 
( - 0.857 pH + 0.494 )  49.29 pH - 60.18 
k CaCO 3 = 10 e 
 R 
 1 1
- 


(1 + x CO 2 ( 0.434 pH - 149 ))
  293 T  

The linear dissolution rate kCaCO3 is calculated in µm s-1, R is the universal gas constant, T is temperature
in K, xCO2 is the mole fraction of CO2 in the gas phase. Ground limestone has a polydisperse
granulometric composition. Based on granulometric analysis, we modelled the size distribution of the
limestone as a fixed set of monodisperse granulometric interval classes covering all particle sizes. In
modeling, the dissolution of limestone particles for each granulometric class was calculated for a certain
time interval. In a model the mixing of various slurry flows with undissolved limestone particles is
performed. In order to simplify these calculations (using fixed particle sizes), the actual contraction of
dissolving particles diameter (mass difference) was transformed to the corresponding number of particles
that fell from their class to a lower size class, considering the mass balance.

Oxidation of sulphite (experimental model).


The main conversion reaction of the absorbed SO2 is oxidation with dissolved oxygen to the SO42- anion:
1
SO3 + O 2 →
2- k ox 2-
SO4
2

The reaction has been studied by many authors, however data are to some extent contradictory. We have
therefore chosen experimentally determined rate for specific systems (flue gas, limestone, water). The
rate of oxidation, kox, was estimated from the SO2 mass balance in the pilot plant at various pH values in
a long-term steady state operation. Well-oxidized conditions were maintained by the dispersion of forced
air to the main slurry conditioning tank. The empirical correlation between the kox and the pH value (at
reaction order 1.5 for sulphite ions), obtained by the linear regression of experimental data, is:

dc
( SO32− )
− dt = kox c(1SO
.5
2−
)
= 10( −1.23 pH +6.62) c(1SO
.5
2−
)
3 3

Precipitation of produced gypsum


Accumulated Ca2+ and SO42- ions precipitate as CaSO4×2H2O. Calcium sulphite can also partially
precipitate as CaSO3×2H2O or CaSO3×1/2H2O or as Ca(SO3)1-x(SO4)x ×0.5H2O, depending on
suspension pH, temperature and dissolved sulphite concentration.

4/18
Ca + SO3 + 2 H 2 O ↔ CaSO 3 × 2 H 2 O( s )
2+ 2-

CaSO 3'
K sp = [ Ca 2+ ] [ SO 32 - ]

Ca + SO3 + 2 H 2 O ↔ CaSO 3 × 2 H 2 O( s )
2+ 2-

CaSO 3'
K sp = [ Ca 2+ ] [ SO 32 - ]

KspCaSO3´ and KspCaSO4´ are solubility products of CaSO3 and CaSO4, respectively, for ideal solutions, and
must be corrected using the activity coefficients for use with non-ideal solutions.

Measurements
Pilot plant with nominal capacity of 14,000 Nm3 h-1 of flue gas was constructed and operated and further
experimental data from test period operation of industrial desulphurisation plant (275 MW) located at
Unit 4, Šoštanj, Slovenia were used to test the developed model of spray tower wet limestone
desulphurization process. Main properties of mentioned pilot and industrial plants are presented in Table
1.

Table 1: Main properties of pilot and industrial desulphurization plant

Value
Property: Pilot plant Industrial plant
Cross section area in the absorber, (m2) 1.3 196
Volume of the main slurry tank, (m3) 30(+30) 5000
Slurry temperature, (°C) 60 60

Slurry pH value, - 5–6 5.65 ± 0.13


Number slurry circulation pumps, - 3 6
Flow rate of each circulation pump, (m3 h-1) 175 max. 7000
Heights of six dispersion nozzles levels, (m) 1.25, 3.4, 5.55, 9.7, 11.2, 12.7,
7.75, 9.85, 12.0 14.3, 15.8, 17.3

5/18
100
95
90
85
80
% of SO2 removal

75
70
pH=5.7 (model)
65
pH=5.5 (model)
60
pH=5.0 (model)
55 pH=5.7 (plant)
50 pH=5.5 (plant)
45 pH=5.0 (plant)

40
0 5 10 15 20 25 30 35 40 45 50 55
L/G (L m-3)

Figure 2: Comparison of pilot plant and modelled dependence of SO2 removal rate (%)
against L/G ratio in the absorber for three pH values of absorbing slurry.

Validation of the model was done on pilot and industrial plant experimental data. An example of
validation of the FGD model with data from pilot plant is presented on Figure 2. The comparison of
experimental and model values of overall SO2 removal rate (%) in the absorber for various L/G ratios in
the absorber and three pH values of input slurry to the absorber is presented. During the test period the
industrial plant operated at conditions described in Table 1 and 2, except for the number of pumps used;
only lower five levels were required in operation. Thus plant reached overall SO2 removal efficiency
97.3 %. Model predicted at mentioned conditions 98.1 % removal efficiency. For additional verification,
the model was tested by variation of other relevant parameters (e.g., L/G ratio, pH value, superficial gas
velocity in absorber).

Optimisation

Using developed model of spray tower absorber, it was possible to study optimisation of its design by
changing the geometry and gas velocity, the height of complete assembly of all six dispersion nozzles
levels and effect of variations in chosen cross section area of the absorber (resulting in different gas
phase superficial velocity uG).
As an example of modelling (required L/G ratio and P against average absorber height) some results are
presented on Figures 3 and 4.

6/18
35.0

30.0

25.0
L/G ratio in absorber (L/m3)

20.0

15.0

10.0

uG= 1.7 m/s


5.0 uG= 2.5 m/s
uG= 4.0 m/s

0.0
0 2 4 6 8 10 12 14 16
average absorber height (m)

Figure 3: Calculated required L/G ratio for spray absorber (for three gas velocities)
against average absorber height.

4.5

4.0

3.5
overall required power, P (MW)

3.0

2.5

2.0

1.5

1.0
uG= 1.7 m/s
uG= 2.5 m/s
0.5
uG= 4.0 m/s

0.0
0 2 4 6 8 10 12 14 16
average absorber height (m)

Figure 4: Calculated overall required power for spray tower (for three gas velocities)
against average absorber height.

It is clear from the presented results, that increasing superficial gas velocity in absorber promotes
absorber performance. This phenomenon is commonly used in novel spray tower designs, and is caused
by increased turbulence around dispersed liquid phase droplets, decreasing both gas and liquid side mass

7/18
transfer resistance. Increased gas velocity results in lower L/G ratio, but also in increased gas phase
pressure drop.

Falling film absorber measurements, modelling and optimisation

Hydrodynamics of the falling film of absorbing slurry and mass transfer from gaseous to liquid phase
In calculations of both mass transfer coefficients for the falling film of liquid various equations from
literature were tested. As it is explained bellow, it was found that calculations for developed wavy flow
regime (ReN < 2000) and for transition regime from developed wavy to turbulent flow (ReN > 2000)
should be separated.
In calculations of the gas side mass transfer coefficient kG, equations from Gilliland (9),
Kafesjian (10) and Spedding and Jones (11) were used. As presented in the results, for liquid film
Reynolds values (ReN) above 2000, the equation from Spedding and Jones (11) gives the best results
according to experiment. The review of results is given on Figure 5.
Falling film absorber was used on laboratory scale apparatus for measurements of SO2 absorption at wet
limestone desulphurisation process conditions. Measurements were used to verify the developed model
of SO2 absorption in falling film absorber consisting of single vertical tube element of given height and
diameter.

85
experimental
80 Higbie/Gilliland
Higbie/Kafesjian
75 Higbie/Spedding
Holpanov/Gilliland
70 Holpanov/Kafesjian
Holpanov/Spedding
Henstock/Gilliland
65 Henstock/Kafesjian
Henstock/Spedding
SO2 removal (%)

60 Davis/Gilliland
Davis/Kafesjian
55 Davis/Spedding
Brauner/Gilliland
50 Brauner/Kafesjian
Brauner/Spedding
45 optimal combination

40

35

30

25

20
0 5 10 15 20 25 30
-3
actual L/G ratio (L m )

Figure 5: Agreement between predicted removal of SO2 for different published models of falling film
hydrology, best combination and experimental data.

8/18
All other relevant sub-models describing chemical equilibrium in liquid phase, dissolution of limestone,
oxidation of sulphite and precipitation of produced gypsum were the same as described for spray
absorber model.

On the Figure 6 very good agreement between experimental and modelled results is shown for the
absorption in a tube with 47 mm diameter, pH 5,6.
4000
3800 mL/m3 SO2, experiment

3500 2000 mL/m3 SO2, experiment


800 mL/m3 SO2, experiment
3000 800 mL/m3 SO2, model
SO2 concentration (mL/m3)

2000 mL/m3 SO2, model


2500 3800 mL/m3 SO2, model

2000

1500

1000

500

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5
position (m)

Figure 6: Experimental and modelled SO2 profile in a vertical channel47 mm in diameter

Optimisation of falling film absorber


Optimisation of single tube element design was performed with the model developed. Required L/G
ratio for the required output SO2 concentration has been calculated. Tube diameters of 28, 35 and 47 mm
were used in calculations. Narrower tubes do not allow higher gas phase velocities (flooding would
occur), wider ones would not provide adequate contact area. Heights of 5, 6 and 7 m were found to be of
main interest. Superficial gas phase velocities of 2.5 and 4.0 m s-1 were considered; slower velocities
would result in rather “wide” absorber design, higher velocities are problematic due to possible flooding
conditions. Examples of the results of optimisation are presented on Figures 7, 8. Power consumptions
for different options have been estimated by means of medelled processes, the results are shown in Table
2.

9/18
60.0
28 mm & 2.5 m/s
35 mm & 2.5 m/s
50.0 47 mm & 2.5 m/s
47 mm & 4.0 m/s
35 mm & 4.0 m/s
L/G ratio in absorber (L/m3)

40.0

30.0

20.0

10.0

0.0
4 4.5 5 5.5 6 6.5 7 7.5
tube absorber height (m)

Figure 7: Calculated required L/G ratio in falling film absorber against tube height,
for 27, 35 and 47 mm tube diameter, and 2.5 and 4.0 m s-1 gas phase velocity.

3.0

2.5
overall required power, P (MW)

2.0

1.5

1.0
28 mm & 2.5 m/s
35 mm & 2.5 m/s
47 mm & 2.5 m/s
0.5
47 mm & 4.0 m/s
35 mm & 4.0 m/s

0.0
4 4.5 5 5.5 6 6.5 7 7.5
tube absorber height (m)

Figure 8: Calculated required overall power P in falling film absorber against tube
height, for 27, 35 and 47 mm tube diameter, and 2.5 and 4.0 m s-1 gas phase velocity.

10/18
Table 2: Overview of optimisation results for two spray tower designs and falling film design.

Spray tower Falling film


Parameter: case 1 case 2
gas flow rate (Nm3 h-1): 1.52 106
uG (m s-1) 2.5 4.0 2.5
L/G ratio (L m-3): 15.5 12 21
pH 5.8
SO2 input (vppm): 2000
SO2 output (vppm): 110
Absorber height /square 17.3/14.6 17.3/11.5 8/17.5
side dimension (m):
PL (MW): 2.367 1.822 1.115
PG (MW): 0.358 0.707 0.14
TOTAL P (MW): 2.725 2.530 1.255
Notes: Scaling potential low Scaling potential
high
Potential of heat
recovery

11/18
Conclusions for falling film absorber
It is evident from presented results of optimisation, that falling film absorber can at approximately equal
L/G ratios reach much lower required power P for the operation (compared to spray tower design). This
is mainly caused by comparably lower height of the absorber itself (e.g. 6 m instead of 12 m), and partial
also by very little pressure drop of gas phase in absorber. Results also show much higher liquid phase
hold up in the absorber and much higher calculated interfacial area in the absorber. Falling film absorber
can thus offer far less expensive operation, much lower height of the absorber, and comparable lateral
dimensions to conventional spray tower design. Further possibility of falling film absorber is the ability
of the removal of the heat from the system and use it for the pre-heating of combustion air or boiler feed
water. The biggest withdrawal of falling film absorber is the increased possibility of formation of
chemical scaling on the surfaces of the internals. Scaling formation rate was therefore investigated in a
continuous pilot arrangement.

SCALING RATE DETERMINATION

The scaling rate was determined in the continuous small pilot FGD arrangement with oven for the
production of hot flue gas doped with the stream of SO2 , absorber, make-up tank with oxidation,
automatic calcite slurry addition for pH control and hydrocyclon for gypsum removal. The method used
was based on the previous work of Zakrajšek (12, 13) which is also similar to method used by Adams
(14). Test plates were prepared from four materials tested. The plates had the diameter of about 50 mm,
with a mounting hole in the middle, and were hanged in the middle of the make-up tank.
The scaling rate was determined by submersing a set of the test plates in a slurry of the tank for a given
time (about one day), with a continuous operation of the FGD apparatus at given experimental
conditions, followed by removal of the set, gentle washing of test plates by distilled water, short drying at
40 °C and weigh of the plates. Normally, the plates were then returned back to the neutralization tank to
continue with next measurement. One measurement per test plate per day was thus obtained. In general,
given set of the experimental conditions were held for a period of 2 - 4 days (with measurements each
day) until consistent data were collected. Experimental work was carried out in a number of campaigns
lasting each up to about two months.
Scaling rate was calculated from weigh change per test plate, its contact area and scaling exposure time
(within a single measurement) and is reported here in mg×m-2×day-1. The results are given on the figure
9. It is evident that different materials behave differently, scaling rate on rubber surfaces is much higher
than on polypropylene and stainless steel. The scaling produced on polypropylene surfaces is not very
compact and can be easily removed by high pressure water spray, as shown on Figure 10.

12/18
2000 20000

15000
1500

10000

Scaling rate for rubber only (mg/(day×m2))


1000
5000
Scaling rate (mg/(day×m2))

500
0

0 SS 5
-5000

SS 7
-10000
-500 PP 6
PP 7
-15000
Net 3
-1000
Rubber 7
-20000
Rubber 8
-1500 Linear (Rubber 8)
-25000
Linear (PP 6)

-2000 Linear (SS 7) -30000


0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016
SO2 loading rate (mol×s-1×m-3)

Figure 9: Scaling formation rate for different materials and different SO2 loading rates

Figure 10: Falling film absorption module made from polypropylene with scaling on the right and after
high pressure water spray cleaning on the left.

13/18
OPTIMISATION OF THE EQUIPMENT LAYOUT BY INTEGRATION OF SUB-PROCESSES
Other field of optimization is the layout of the equipment and the integration of sub-processes in the
same equipment. Some integration steps proposed, designed and tested are as follows, shown in figures
11 and 12 with investment and operating costs estimates given in tables 3 and 4:
• Intermediate product (gypsum) storage is done in the absorber make-up tank, by allowing the
density to rise up to 20 % solids.
• Calcite slurry is instantaneously produced from calcite powder wetted by absorption slurry.
• Hydrocyclons are fed directly from re-circulating pumps and underflow is fed directly to the
automatic filter or centrifuge.

Figure 11: Scheme of standard wet limestone process

14/18
Figure 12: Scheme of optimized wet limestone process, case with falling film absorption

15/18
Table 3: Summary of the results for low cost FGD technology applied for 215 MW thermal power plant.
Four alternative cases are considered from the corresponding commercial offers.

Alternative cases for FGD plant considered


SPRAY FALL.FILM
Case name: SPRAY FALL.FILM
+GAVO +GAVO
Case: A B C D
No. Item Unit Values
1 Power plant output MWe 215 215 215 215
2 Flue gas input flow rate, wet Nm3/h 802,031 802,031 802,031 802,031
3 Flue gas output flow rate, sat. wet Nm3/h 876,639 876,639 876,639 876,639
Flue gas SO2 input concentration,
5 dry, 6% O2 mg/Nm3 5,727 5,727 5,727 5,727
Flue gas SO2 output conc., dry,
6 6%O2 mg/Nm3 360 360 360 360
7 Flux of SO2 absorbed t/h 3.61 3.61 3.61 3.61
8 Calcite purity % 97.0 97.0 97.0 97.0
9 Calcite efficiency % 97.0 97.0 97.0 97.0
10 Calcite consumption t/h 5.99 5.99 5.99 5.99
11 Gypsum production, dry t/h 10.49 10.49 10.49 10.49
12 Gypsum production, wet t/h 12.2 12.2 12.2 12.2
13 Gypsum moisture % 14.0 14.0 14.0 14.0
14 Number of recircul. pumps - 4 4 4 4
15 Liquid flow total m3/t 16,218 16,218 (7890) + 14.903 (7890) + 14.903
16 Solids in suspension % 0 - 20 10 - 20 10 - 20 10 - 20
17 Suspension density, max kg/m3 1,089 1,089 1,089 1,089
18 Average hydrostatic elevation m 10.40 10.40 (2,0) + 8,50 (2,0) + 8,50
19 Pressure drop on nozzle bar 0.60 0.60 0.01 0.01
20 Total water consumption t/h 65 65 65 65
21 Total power for pumps kW 952 952 486 486
22 Pressure drop for gas in FGD mbar 21 12 14 5
23 Additional power for fan kW 650 378 427 155
24 Other power consumers kW 399 309 399 309
25 Total power kW 2.00 1.64 1.46 1.15
26 Calcite price Euro/t 22.0 22.0 22.0 22.0
27 Electrical power price Euro/kWh 0.042 0.042 0.042 0.042
28 Water price Euro/t 0.042 0.042 0.042 0.042
29 Add. costs due to FGD - calcite Euro/t SO2 36.5 36.5 36.5 36.5
30 Add. costs due to FGD - electricity Euro/t SO2 23.1 18.9 16.8 13.3
31 Add. costs due to FGD - water Euro/t SO2 0.8 0.8 0.8 0.8
32 Add. costs for consumables in FGD Euro/t SO2 60.4 56.2 54.1 50.6
33 Costs for maintenance in FGD Euro/t SO2 6.6 6.2 6.2 5.7
34 Costs for labor in FGD Euro/t SO2 2.6 2.6 2.6 2.6
35 Investment total: Million Euro 15.8 14.7 14.8 13.6
36 Specific investment: Euro/kWe 73.3 68.4 68.9 63.0
37 Specific investment: US$/kWe 91.6 85.5 86.1 78.8

16/18
Table 4: Comparison of the specific construction costs for four alternative low cost FGD plants
(LCFGD) for TPP Tu case, against three base line (classic) FGD plants constructed or under
construction.

SPECIFIC INVESTMENT
Flue gas Flue gas
TYPE (CASE) / YEAR Power Power SO2 input SO2 output
No. US$/kW % % %
OF START UP plant (MWe) conc., conc.,
mg/m3 mg/m3
1 CLASSIC / 1994 S4 275 6,500 400 336 100
2 CLASSIC / 2000 S5 325 6,500 400 308 100 91
CLASSIC - OPTIMIZED
3 / 2004 T 125 10,000 800 142 100 46 42
LCFGD (SPRAY +
4 GAVO) / 2004 Tu6 215 5,700 400 92 64 30 27
5 LCFGD (SPRAY) / 2004 Tu6 215 5,700 400 85 60 28 25
LCFGD (FALL.FILM +
6 GAVO) / 2004 Tu6 215 5,700 400 86 61 28 26
LCFGD (FALL.FILM) /
7 2004 Tu6 215 5,700 400 79 55 26 23

Boundary conditions and other data used in above table:


Item Unit Value
Maintenance cost per year in % of invest. % 1
Number of workers for FGD plant operation - 6
Bruto worker salary Euro/month 850
US$/Euro exchange rate: US$/Euro 1.25

Detailed case study was prepared for a 215 MW PP. Up to 30 % lower investments can be expected in
comparison with current offers on the market. In respect to operating costs, the energy consumption and
cost of maintenance could be decreased by 10 to 25%. As shown by risk analysis the availability of the
system remains equally high as for classical lay-outs.

CONCLUSIONS

Rigorous no-stationary models have been developed for spray absorption and falling film absorption in
wet limestone FGD process. Models have been experimentally verified and used for optimisation of case
FGD projects. At the present state of development the models consider realistic chemistry and limestone
dissolution but simplify the pattern of the streaming of gaseous phase, assuming the uniform gas velocity
distribution across the absorption device. In this simplification we see a significant potential for the
improvement of the models: these physical-chemical models should be combined with the realistic
model for aerodynamics of the gaseous phase streaming in different geometries of the absorption
equipment.
New facts have been experimentally developed that allow the consideration of falling film absorbers that
could avoid scaling problems in long-term operation.
The integration of the equipment leads to the reduction of some tanks, valves, pumps, pipes and required
space and thus offers significant potential for investment and operating savings. Diminished number of
the equipment in use decreases also the risk of the failure of the whole system, as shown by risk analysis.

17/18
REFERENCES

(1) Kronig, R. and Brink, J.C., 1950, On the theory of extraction from falling droplets, Appl. sci. Res.
(2) Stergaršek A., Kocjančič, R., Gerbec M., Chem. Eng. Sci., (1996) 51, 23, pp. 5081 - 5089.
(3) Smolik J. and Vitovec J., J. Aerosol Sci., (1984) 15, pp. 545-552
(4) Carmichael G.R., Peters L.K., Atmos. Env., (1979) 13, pp. 1505 - 1513.
(5) Brewer, L., Thermodynamic values for desulphurization processes, in Hudson, J.L. and Rochelle,
G.T., editors, 1982, Flue gas desulfurization, ACS symposium series, Washington D.C., 188, 1-39.
(6) Luckas M., Lucas K., Roth H., AIChE J., (1994) 40, No. 11, pp. 1892 - 1900
(7) Gerbec M., Stergaršek, A., Kocjančič, R., Acta Chimica Slovenica, (1990) 37, No. 3., pp. 279 -
197.
(8) M. Gerbec, A. Stergaršek, R. Kocjančič, Comput. Chem. Eng. 1995, 19, Suppl., S283 -S286.
(9) Gilliland E.R., Sherwood T.K., Ind. Eng. Chem., (1934) 26, pp. 516.
(10) Kafesjian R., Plank C. A., Gerhard E. R., AIChE J, (1961) 7, No. 3, pp. 463 - 466.
(11) Spedding P.L., Jones M.T., Chem. Eng. J., (1988) 37, pp. 165 - 176.
(12) J. Nývlt, J. Maček, S. Zakrajšek, Cryst. Res. Technol. 1993, 28(4), 479-486.
(13) J. Maček, S. Zakrajšek, J. Radikovič, V. Bole, J. Cryst. Growth 1993, 132, 99-10
(14) J.F. Adams, V.G. Papangelakis, Can. Metall. Q. 2000, 39(4), 421-432.

18/18

Você também pode gostar